The high-temperature Metamorphic Core Complex (, ): Geometry, kinematics and thermal structure Alexandre Beaudoin, Romain Augier, Valentin Laurent, Laurent Jolivet, Abdeltif Lahfid, Valérie Bosse, Laurent Arbaret, Aurélien Rabillard, Armel Menant

To cite this version:

Alexandre Beaudoin, Romain Augier, Valentin Laurent, Laurent Jolivet, Abdeltif Lahfid, et al.. The Ikaria high-temperature Metamorphic Core Complex (Cyclades, Greece): Geometry, kinematics and thermal structure. Journal of Geodynamics, Elsevier, 2015, 92, pp.18-41. ￿10.1016/j.jog.2015.09.004￿. ￿insu-01225362￿

HAL Id: insu-01225362 https://hal-insu.archives-ouvertes.fr/insu-01225362 Submitted on 6 Nov 2015

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés.

Distributed under a Creative Commons Attribution - NoDerivatives| 4.0 International License Accepted Manuscript

Title: The Ikaria high-temperature Metamorphic Core Complex (Cyclades, Greece): Geometry, kinematics and thermal structure

Author: Alexandre Beaudoin Romain Augier Valentin Laurent Laurent Jolivet Abdeltif Lahfid Valerie´ Bosse Laurent Arbaret Aurelien´ Rabillard Armel Menant

PII: S0264-3707(15)30025-9 DOI: http://dx.doi.org/doi:10.1016/j.jog.2015.09.004 Reference: GEOD 1390

To appear in: Journal of Geodynamics

Received date: 29-1-2015 Revised date: 18-9-2015 Accepted date: 28-9-2015

Please cite this article as: Beaudoin, A., Augier, R., Laurent, V., Jolivet, L., Lahfid, A., Bosse, V., Arbaret, L., Rabillard, A., Menant, A.,The Ikaria high-temperature Metamorphic Core Complex (Cyclades, Greece): Geometry, kinematics and thermal structure, Journal of Geodynamics (2015), http://dx.doi.org/10.1016/j.jog.2015.09.004

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. *Highlights (for review)

Ikaria Island corresponds to a large-scale migmatite-cored MCC.

Thermal structure revealed by RSCM shows a drastic increase from top to bottom.

The MCC was exhumed by two detachments through the brittle-ductile transition.

Migmatites were dated to 15.7 ± 2 Ma by U-Th-Pb analysis on monazites.

A large-scale high temperature zone is proposed for the central part of the Aegean.

Accepted Manuscript

Page 1 of 86 *Manuscript Click here to view linked References

1 The Ikaria high-temperature Metamorphic Core Complex (Cyclades, Greece): Geometry,

2 kinematics and thermal structure.

3

4 Alexandre Beaudoina,b,c,*, Romain Augiera,b,c, Valentin Laurenta,b,c, Laurent Joliveta,b,c,

5 Abdeltif Lahfidb,c, Valérie Bossed, Laurent Arbareta,b,c, Aurélien Rabillarda,b,c, Armel

6 Menanta,b,c

7

8 aUniversité d’Orléans, ISTO, UMR 7327, 45071, Orléans, France

9 bCNRS/INSU, ISTO, UMR 7327, 45071 Orléans, France

10 cBRGM, ISTO, UMR 7327, 45060 Orléans, France

11 dUniversité Blaise Pascal, Laboratoire Magmas & Volcans, CNRS, UMR 6524, 63000,

12 Clermont-Ferrand, France

13

14 *Corresponding author: Alexandre Beaudoin, ISTO, 1A rue de la Férollerie, 45071, Orléans,

15 France (e-mail: [email protected]; Tel: +33 2 38 49 25 73)

16

17

18 Abstract

19

20 This work attempted at clarifying the structure of Ikaria using primarily intensive

21 geological mappingAccepted combined with structural analysis Manuscript and a geothermometry approach of

22 Raman spectrometry of carbonaceous material. Foliation over the whole island defines a

23 structural dome cored by high-grade to partially-molten rocks. Its exhumation was completed

24 by two top-to-the-N ductile extensional shear zones, operating in the ductile and then the

25 brittle fields, through a single extensional event coeval with progressive strain localization.

Page 2 of 86 26 The thermal structure of the dome with regard to position of ductile shear zones was retrieved

27 using the Raman spectroscopy of carbonaceous material. Peak-metamorphic temperatures

28 range from 390 °C in the upper parts of the structure down to 625 °C in the core of the dome

29 in the vicinity of migmatites and S-type granite. Pioneer in situ U-Th-Pb analyses on monazite

30 performed on the leucosome parts of these rock yielded a 15.7 ± 0.2 Ma age. Ikaria Island

31 thus completes the series of Miocene migmatite-cored Metamorphic Core Complex in the

32 central part of the Aegean domain where a genuine high-temperature zone can be defined as

33 the central Aegean HT zone. There, the extreme stretching of the continental crust is

34 associated with dominantly top-to-the-N kinematics.

35

36 Keywords: Structural analysis; RSCM geothermometry; U-Th-Pb geochronology;

37 Metamorphic Core Complex; Ikaria; North Cycladic Detachment System.

38

39

40 1. Introduction

41

42 In the Mediterranean realm, the retreat of oceanic slabs triggered the initiation of back-

43 arc extension (i.e large-scale extension in the upper plate of a subduction zone), leading to the

44 collapse of previously thickened continental lithosphere (Le Pichon and Angelier, 1979;

45 Malinverno and Ryan, 1986; Dewey, 1988; Platt and Vissers, 1989; Royden, 1993; Jolivet

46 and Faccenna, 2000;Accepted Rosenbaum et al., 2002; Faccenna Manuscript et al., 2004; Jolivet et al., 2008). This

47 post-orogenic evolution (i.e. crustal thinning by extension-related normal faulting after an

48 episode of crustal thickening) resulted in the formation of series of extensional domains or

49 wide-rift systems (Lister et al., 1984; Buick, 1991; Corti et al., 2003) such as the Alboran Sea,

50 the Tyrrhenian Sea, the Pannonian Basin and the . Lateral evolution from the

Page 3 of 86 51 central parts of the extensional domains to the bounding non-collapsed orogenic segments

52 implies drastic lateral gradients of finite extension and suggests highly non-cylindrical

53 structures. In the Aegean Sea, along with the drastic decrease in topography and crustal

54 thickness, the main orogenic structures of the Hellenic belt are increasingly reworked by

55 extension from continental Greece to , in the center of the Aegean Sea. Extensional

56 structures related to back-arc crustal stretching evolve from essentially brittle steep and

57 shallow-dipping normal faults in continental Greece to shallow-dipping ductile shear zones in

58 the center of the Aegean domain (Jolivet and Patriat, 1999; Jolivet et al., 2010). Furthermore,

59 a straightforward correlation between the degree of non-coaxiality of the back-arc domain and

60 the amount of stretching of the continental crust is observed at the scale of the entire Aegean

61 domain (e.g. Augier et al., 2015). Marginal areas that connect to non-collapsed orogenic

62 segments display symmetrically arranged detachment systems. In the western parts of the

63 Aegean domain, both the West Cycladic Detachment System (WCDS) and the North Cycladic

64 Detachment System (NCDS) exhume a horst-shaped domain, where orogenic features are still

65 nicely preserved, depicting bivergent extension (e.g. Jolivet et al., 2010; Grasemann et al.,

66 2012). Conversely, in the center of the Aegean domain, deformation remains highly

67 asymmetric from Mykonos in the north, all the way to Sikinos in the south (e.g. Gautier et al.,

68 1993; Kumerics et al., 2005; Denèle et al., 2011; Augier et al., 2015). There, orogenic features

69 are particularly overprinted or even locally erased by the combined effects of intense top-to-

70 the-N shearing and partial-melting. Migmatite-cored Metamorphic Core Complexes (MCCs),

71 associated with AcceptedMiocene intrusions, roofed by major Manuscript top-to-the-N crustal-scale detachments,

72 are described on Naxos, in the center (e.g. Lister et al., 1984; Urai et al., 1990; Buick, 1991;

73 Gautier and Brun, 1994; Jolivet et al., 2004a; Vanderhaeghe, 2004) and Mykonos in the north

74 of the Aegean domain (Lecomte et al., 2010; Denèle et al., 2011). Concentrating a large part

75 of the total amount of stretching, recognition of MCCs therefore appears of prime importance.

Page 4 of 86 76 Similarly, asymmetry of crustal thinning at the regional-scale is another key-question for

77 understanding back-arc extension dynamics. However, the current understanding of back-arc

78 dynamics in the Aegean domain is hindered by the severe lack of knowledge for the bulk of

79 its eastern part.

80 One of the largest Aegean islands, displaying the largest intrusion of the Aegean Sea,

81 located between the northern Cyclades and western Turkey, Ikaria Island has been the focus

82 of several recent studies. However, the first order structural architecture of this island remains

83 conflicting and the existing geological maps of Ikaria present marked discrepancies. Besides,

84 the current knowledge of the metamorphic record and particularly the thermal structure of

85 Ikaria remain fragmentary. These problems were reconsidered after an extensive field survey,

86 including primarily new geological mapping and structural analysis. The position and

87 importance of the various tectonic contacts as well as the bulk thermal architecture of the

88 island were further constrained using the geothermometry approach of Raman Spectrometry

89 of Carbonaceous Material (RSCM). Following their recent discovery, the migmatites

90 described in the core of the structure were dated by U-Th-Pb LA-ICPMS analyses on

91 monazite.

92

93

94 2. Geological setting

95

96 2.1. GeodynamicAccepted context Manuscript

97

98 The Aegean domain (Fig. 1) corresponds to the collapsed segment of the Hellenides-

99 Taurides belt, developed as the result of the convergence between Apulian and European

100 plates, in the eastern Mediterranean, since the Late Cretaceous (e.g. Aubouin and Dercourt,

Page 5 of 86 101 1965; Brunn et al., 1976; Bonneau and Kienast, 1982; Bonneau, 1984; van Hinsbergen et al.,

102 2005; Jolivet and Brun, 2010, Ring et al., 2010). During this period, a south-verging crustal-

103 scale orogenic wedge was formed by subduction and accretion of several Apulia-derived

104 continental blocks separated by oceanic basins. Post-orogenic extension of the Hellenic

105 thickened crust started in the Early Oligocene (Jolivet and Faccenna 2000; Jolivet et al., 2008)

106 or the Early Miocene (Ring et al., 2010) by a combination of gravitational collapse and back-

107 arc extension during the southward retreat of the African slab (e.g. Jolivet and Faccenna,

108 2000; Jolivet and Brun, 2010). Intense crustal stretching of the upper plate leads to the

109 formation of a series of MCCs (Lister et al., 1984; Avigad and Garfunkel, 1989; 1991;

110 Gautier and Brun, 1994; Jolivet et al., 2004a). Despite the importance of the post-orogenic

111 overprint on the orogenic architecture, the original vertical superposition of tectonic units in

112 the nappe stack is often preserved at the scale of the whole Aegean domain (e.g. Bonneau,

113 1984; van Hinsbergen et al., 2005). Three main tectonometamorphic units are classically

114 recognized (e.g. Bonneau, 1984; Ring et al., 2010) (Fig. 1):

115 1) The Upper Cycladic unit corresponds to a lateral equivalent of the Pelagonian nappe

116 (e.g. Bonneau, 1984; Jolivet et al., 2004a) recognized in continental Greece. This unit

117 generally crops out as isolated klippes or rafts in the Cyclades, essentially made of ophiolitic

118 material such as in , , Mykonos, Kea, Kythnos, Serifos and (e.g. Ring et

119 al., 1999; Jolivet et al., 2010; Grasemann et al., 2012). Rocks preserve a Cretaceous HT-LP

120 metamorphic imprint but escaped both the Eocene HP-LT and the Oligocene-Miocene HT-LP

121 tectonometamorphicAccepted events (Katzir et al., 1996). Manuscript Syn-tectonic detrital shallow-marine and

122 continental sediments locally form the uppermost unit on Mykonos, , Naxos and Ikaria

123 (e.g. Angelier, 1976; Photiades, 2002a; Sánchez-Gómez et al., 2002; Kuhlemann et al., 2004;

124 Lecomte et al., 2010). Conglomerates mostly contain pebbles derived from the Upper

Page 6 of 86 125 Cycladic nappe and reworked magmatic rocks as young as 10 Ma (e.g. Sánchez-Gómez et al.,

126 2002).

127 2) The Cycladic Blueschists unit crops out as a composite unit including locally a

128 significant component of metabasic rocks interleaved with metapelites and marbles, all

129 equilibrated in blueschist-facies conditions (e.g. Blake et al., 1981; Bonneau, 1984; Avigad

130 and Garfunkel, 1991; Keiter et al., 2004). This unit experienced a complex alpine

131 tectonometamorphic evolution, with an early burial in HP-LT conditions reaching blueschist

132 to eclogite-facies conditions during the Eocene, followed by a greenschist overprint of

133 variable intensity during the Oligocene and the Miocene (Altherr et al., 1979, 1982; Wijbrans

134 et al., 1990; Parra et al., 2002; Duchêne et al., 2006; Augier et al., 2015). On Ikaria, despite

135 the lack of reliable HP index minerals, the Messaria unit (Kumerics et al., 2005) was

136 correlated with the Ampelos nappe recognized on Samos that experienced metamorphic

137 conditions of the order of 15 kbar and 500 °C (Will et al., 1998).

138 3) The lower units crop out in tectonic windows. The Cycladic Basement unit crops

139 out in the central and southern Cyclades (i.e. on Naxos, Paros, Sikinos and Ios) (e.g.

140 Andriessen et al., 1987). It is composed of Variscan granitoids mantled by micaschists that

141 retain either metamorphic relics of amphibolite-facies assemblages or inherited radiometric

142 ages suggesting a complex prealpine history (e.g. Henjes-Kunst and Kreuzer, 1982;

143 Andriessen et al., 1987; Keay and Lister, 2002). It is sometimes covered by Mesozoic marbles

144 that may represent a HP equivalent of the Gavrovo unit cropping out in continental Greece

145 (Jolivet et al., 200Accepted4b). Just as the Cycladic Blueschists Manuscript unit, the Cycladic Basement unit

146 shows a complex alpine tectonometamorphic evolution, with an initial subduction-related

147 burial in HP-LT conditions during the Eocene whose trace has been obscured by an

148 Oligocene-Miocene local overprint (van der Maar et al., 1981; Vandenberg and Lister, 1996;

149 Baldwin and Lister, 1998; Augier et al., 2015). On Naxos and Paros, it experienced a partial-

Page 7 of 86 150 melting stage in amphibolite to granulite-facies conditions (i.e. Jansen and Schuiling, 1976;

151 Buick and Holland, 1989; Vanderhaeghe, 2004; Duchêne et al., 2006).

152 Rocks of the Cycladic Blueschists and the lower units were exhumed during two

153 distinctive stages in two contrasted geodynamic settings. The first stage occurred in the

154 Hellenic subduction context, during the Eocene, with burial and synorogenic exhumation (by

155 shortening-related normal faulting) of blueschist to eclogite-facies assemblages in an

156 extrusion wedge structure (e.g. Altherr et al., 1979; Wijbrans et al., 1990; Trotet et al., 2001a;

157 2001b; Groppo et al., 2009; Ring et al., 2007; Jolivet and Brun, 2010). The second stage

158 occurred in the Oligocene-Miocene with the post-orogenic exhumation in the back-arc

159 domain of the Cycladic Blueschists and the lowers units as metamorphic domes or migmatite-

160 cored MCCs (Lister et al., 1984) below a series of detachments (e.g. Avigad and Garfunkel,

161 1989; Buick and Holland, 1989; Buick, 1991; Faure et al., 1991; Lee and Lister, 1992;

162 Gautier et al., 1993; Gautier and Brun, 1994; Jolivet and Patriat, 1999; Keay et al., 2001;

163 Vanderhaeghe, 2004; Kumerics et al., 2005; Mehl et al., 2005; 2007; Denèle et al., 2011;

164 Augier et al., 2015). In the northern Cyclades, these detachments were recently grouped in a

165 single large-scale top-to-the-N structure running over 130 km to form the NCDS (Fig. 1;

166 Jolivet et al., 2010). A similar set of detachments with opposed, top-to-the-S or SW

167 kinematics, was identified in the western Cyclades (Grasemann and Petrakakis, 2007; Iglseder

168 et al., 2009, 2011; Tschegg and Grasemann, 2009; Brichau et al., 2010) and recently

169 mechanically linked to form the WCDS (Grasemann et al., 2012). The Naxos-Paros

170 Detachment (NDP)Accepted completes those series of detachment Manuscript systems in the center of the Aegean

171 domain (Buick and Holland, 1989; Buick, 1991; Gautier and Brun, 1994; Kruckenberg et al.,

172 2011). Another major shear zone over which top-to-the-N kinematics rework top-to-the-S

173 shear sense occurs in the southern Cyclades on the islands of Sikinos and Ios (e.g.

174 Vandenberg and Lister, 1996; Forster and Lister, 2009; Huet et al., 2009; Thomson et al.,

Page 8 of 86 175 2009; Augier et al., 2015). Extension on this shear zone starts to exhume rocks near the Early

176 Miocene (Thomson et al., 2009). Here, top-to-the-S kinematics was first considered as

177 extensional (Vandenberg and Lister, 1996; Forster and Lister, 2009; Thomson et al., 2009)

178 and later correlated with the WCDS (Ring et al., 2011). Another study reinterpreted this shear

179 zone as a top-to-the-S thrust reworked by top-to-the-N extension (the South Cycladic Thrust,

180 SCT) (Huet et al., 2009). The same interpretation is made on Sikinos where the SCT crops out

181 (Augier et al., 2015). Late exhumation stages along extensional systems were accompanied by

182 the emplacement of syn-tectonic Miocene I and S-type granites (i.e. Tinos, Mykonos, Naxos,

183 Serifos and Ikaria; Fig. 1) (Faure et al., 1991; Lee and Lister, 1992; Altherr and Siebel, 2002;

184 Grasemann and Petrakakis, 2007; Ring, 2007; Iglseder et al., 2009; Bolhar et al., 2010;

185 Laurent et al., 2015). The activity of detachments is also constrained by syn-tectonic

186 deposition of sediments over the Upper Cycladic unit (e.g. Sanchez-Gomez et al., 2002;

187 Kuhlemann et al., 2004; Lecomte et al., 2010; Menant et al., 2013).

188 Although part of these features of the Cycladic geology are found further east within

189 the Menderes massif (Fig. 1), type, distribution and timing of metamorphism appear less

190 clear. The Cycladic Blueschists unit and the Lycian Nappe that experienced HP-LT

191 metamorphic conditions rest on top of the structure to the north and to the south of the massif

192 (Oberhänsli et al., 1998; Rimmelé et al., 2003a; 2003b; Pourteau et al., 2010). HP-LT

193 metamorphic conditions are dated from Late Cretaceous to Eocene (Oberhänsli et al., 1998;

194 Pourteau et al., 2013). Crustal stretching responsible for the final exhumation of the

195 metamorphic partAccepted of the Menderes massif under shallow Manuscript-dipping shear zones (e.g. Rimmelé et

196 al., 2003b; Bozkurt, 2007; van Hinsbergen; 2010; Bozkurt et al., 2011) is very similar to the

197 Cyclades one. Detachments juxtapose Neogene syn-tectonic sediments on metamorphic rocks

198 that are strongly retrograded in the greenschist facies during the Miocene (Hetzel et al.,

Page 9 of 86 199 1995a; Lips et al., 2001). Here again, crustal thinning is accompanied by the emplacement of

200 syn-tectonic granites (Ring and Collins, 2005; Glodny and Hetzel, 2007).

201

202

203 2.2. Geology of Ikaria

204

205 Ikaria is a 40 km-long island located in the eastern Aegean Sea between Mykonos and

206 Samos. A recent map coverage was performed by the Greek Institute of Geology and Mineral

207 Exploration (G-IGME) (Photiades, 2002b), complemented by independent geological maps

208 presenting highly conflicting interpretations (Papanikolaou, 1978; Kumerics et al., 2005;

209 Ring, 2007; Bolhar et al., 2010; Kokkalas and Aydin, 2013).

210 At first glance, the geology of Ikaria consists in an equal distribution of a metamorphic

211 domain to the east and a large-scale magmatic complex to the west, bounded by sedimentary

212 rocks (Fig. 2). The metamorphic part consists in a 1500 m-thick tectonometamorphic

213 succession made of metasediments including metapelites, metaquartzites, marbles and minor

214 metabasites occurrences passing upward to finely alternating metapelites and marbles

215 (Photiades, 2002a; 2002b; Kumerics et al., 2005). The metamorphic grade decreases upward

216 from widespread amphibolite-facies associations (i.e. staurolite-garnet-biotite in metapelites

217 and hornblende-plagioclase in metabasites) to greenschists-facies associations in the

218 uppermost parts of the succession (e.g. Altherr et al., 1982; Kumerics et al., 2005; Martin et

219 al., 2011). PeakAccepted-metamorphic conditions, as retrieved Manuscript from pseudo-section approaches,

220 yielded 6-8 kbar for 600-650 °C conditions for the basal parts of the tectonometamorphic

221 succession (Kumerics et al., 2005; Martin et al., 2011), fringing partial-melting conditions

222 assuming water saturation conditions (Weinberg and Hasalova, 2015). While similar

223 quantitative P-T estimates are currently lacking for the upper parts of the succession, a

Page 10 of 86 224 localized change in the metamorphic grade has however been proposed within the upper parts

225 (Papanikolaou, 1978; Altherr et al., 1982; Kumerics et al., 2005). Besides, it is noteworthy

226 that traces of an initial HP imprint is currently lacking on Ikaria, at variance with neighboring

227 islands (Altherr et al., 1982; Photiades, 2002a; Kumerics et al., 2005; Ring, 2007).

228 Three main magmatic intrusions were recognized on Ikaria (Papanikolaou, 1978;

229 Photiades, 2002b; Ring, 2007); two small-scale, less than 10 km² S-type two-mica

230 leucogranite intrusions (i.e. the Xylosyrtis and the Karkinagrion intrusions) and a large-scale

231 I-type intrusion (i.e. the intrusion). Along with a related pervasive pegmatite dyke

232 array (Photiades, 2002b; Hezel et al., 2011), the Xylosyrtis pluton displays a clear intrusive

233 character within the metamorphic series. Conversely, the nature of the Raches granite contact

234 remains highly controversial and was mapped so far as intrusive (Papanikolaou, 1978;

235 Laurent et al., 2015), as a detachment (Kumerics et al., 2005) or as a thrust (Photiades, 2002a;

236 2002b; Kokkalas and Aydin, 2013). Emplacement of the main intrusions occurred in a narrow

237 15-13 Ma age-range (Bolhar et al., 2010).

238 Metamorphic and intrusive rocks experienced an intense top-to-the-N shearing in both

239 ductile and then brittle regimes (Kumerics et al., 2005; Ring, 2007). The Raches intrusion in

240 the west thus shows a top-to-the-N strain gradient toward the upper structural levels, from

241 proto-mylonites to ultra-mylonites, and finally cataclasites (Laurent et al., 2015). Prior to this

242 study, two main tectonic contacts, along which deformation is concentrated, were

243 distinguished: the Messaria and Fanari detachments (Kumerics et al., 2005). According to

244 Kumerics et al. Accepted(2005), the Messaria detachment correspManuscriptonds to a mylonite zone later partly

245 overprinted by cataclastic deformation. At variance, the Fanari detachment is currently

246 regarded as a purely brittle contact roofing the metamorphic rocks and bounding the

247 sediments of the upper unit (Kumerics et al., 2005; Ring, 2007). This unit, regionally known

248 as the Fanari unit, consists in sandstones, siltites and conglomerates containing clasts of red

Page 11 of 86 249 cherts and ophiolitic rocks of Early Cretaceous age and large-scale olistoliths of Triassic

250 recrystallized limestones (Papanikolaou, 1978) reminding typically rocks of Pelagonian

251 affinity (Pe-Piper and Photiades, 2006). Ages of those sediments spread from Oligocene to

252 Pliocene (Photiades, 2002a; 2002b). In the center of the island, the recrystallized limestone of

253 Kefala, inferred as Triassic, is sometimes described as a klippe of the upper unit of Ikaria

254 (Papanikolaou, 1978; Photiades, 2002a; 2002b; Pe-piper and Photiades, 2006). The main

255 argument for the presence of the Pelagonian unit relies on the description of a diorite intrusive

256 body that yielded Cretaceous K/Ar ages on hornblende (Altherr et al., 1994).

257 Precise age-constraints on peak-metamorphic conditions and the accurate timing for

258 the onset of extensional motions along main shear zones are currently lacking on Ikaria. First

259 record of cooling, below 550 °C (Villa, 1998), is scattered between 25 and 17 Ma (K/Ar on

260 hornblende; Altherr et al., 1982). Late exhumation stages are, in turn, well constrained after

261 15 Ma by varied thermochronological tools (Altherr et al., 1982; Kumerics et al., 2005). The

262 K/Ar and Ar/Ar ages on both fabric-forming white mica and biotite yield numerous 12 to 9

263 Ma ages. In metamorphic rocks, these ages are interpreted as both cooling or deformation

264 ages during exhumation while they are considered as cooling ages for granites (Altherr et al.,

265 1982; Kumerics et al., 2005). Fission-track (FT) analyses yielded 10.3 ± 0.3 to 7.1 ± 0.3 Ma

266 ages for zircon and 8.4 ± 0.8 to 5.9 ± 0.8 Ma ages for apatite (Kumerics et al., 2005). Final

267 cooling stages were constrained by U-Th/He analyses on apatite that yielded symmetrically

268 arranged 6 Ma ages for both flanks and a 3 Ma age for the core of the structure (Kumerics et

269 al., 2005). CoolingAccepted rates therefore exceeded 100 °C/MaManuscript from the ductile to brittle transition

270 (i.e. 300-400 °C, Stöckhert et al., 1999; Imber et al., 2001) and temperatures as low as 70 °C

271 between 10 to 6 Ma.

272

273

Page 12 of 86 274 3. A new geological map of Ikaria

275

276 An extensive field survey was carried out on Ikaria including new field mapping in

277 order to complement the existing G-IGME geological map (Photiades, 2002b). The result is

278 presented in Fig. 2; readers are referred to existing maps showing very conflicting

279 interpretations to appreciate the modifications (Papanikolaou, 1978; Photiades, 2002b;

280 Kumerics et al., 2005; Ring, 2007; Bolhar et al., 2010; Kokkalas and Aydin, 2013).

281 For clarity, the lithostratigraphic subdivisions of Photiades (2002b) were kept as much

282 as possible. Three main lithologies, including marbles, micaschists (including calcschists and

283 minor metaquartzites and metabasites occurrences) and granitic rocks derived from three main

284 intrusive bodies were mapped (Fig. 2). Definition of the main tectonic units changed as the

285 result of the modification and the reinterpretation of their boundaries, particularly the tectonic

286 and intrusive contacts. Three main tectonic units, from bottom to top: the Ikaria, Agios-

287 Kirykos and Fanari units, limited by two major shear zones, are now distinguished. The main

288 tectonic features having a map-scale expression were also reported in detail. These structures

289 consist primarily in the Agios-Kirykos and the Fanari ductile shear zones developed as ramp-

290 flat extensional structures either at shallow-angle to the compositional layering as décollement

291 zones or cutting down-section. The Fanari shear zone even presents evidence for subsequent

292 displacements in the brittle field superimposed on ductile features. On the map, the Fanari

293 detachment clearly cuts across the whole Agios-Kirykos unit and the Agios-Kirykos shear

294 zone. Other shearAccepted zones or brittle contacts put forward Manuscript on previous maps were abandoned.

295 The basal contact below Kefala marble (e.g. Papanikolaou, 1978; Altherr et al., 1994;

296 Photiades, 2002a; 2002b; Pe-piper and Photiades, 2006) is considered less important than in

297 previous studies since no metamorphic gap is backed up by published data. Moreover,

298 detailed mapping of this zone indicate that the diorite is intrusive in Ikaria unit. However, a

Page 13 of 86 299 panorama on the Kefala marble seen suggests the presence of a tectonic contact at its base

300 (Papanikolaou, 1978). Although probably minor, this contact might be a late brittle expression

301 of the system of detachments. Similarly, the nature of the eastern contact of the Raches

302 granite considered either as the lateral equivalent of the Messaria detachment (Kumerics et al.,

303 2005; Ring, 2007) or a major thrust contact (Photiades, 2002a; 2002b; Kokkalas and Aydin,

304 2013) was re-evaluated. Field work in the vicinity of the contact of the Raches granite

305 unambiguously shows, despite ductile deformation, the clear intrusive character of the granite

306 within Ikaria unit (Fig. 3a) as suggested in earlier studies (e.g. Papanikolaou, 1978; Laurent et

307 al., 2015). Along with the Fanari detachment, the Fanari sedimentary unit of Kumerics et al.

308 (2005) was extended further southwest (Fig. 3b). This sedimentary unit was correlated to

309 sediments recognized on the northern part of Ikaria at Gialiskari (Fig. 3c) as initially proposed

310 by Photiades (2002a; 2002b) and recently reinforced by Laurent et al. (2015). There, the

311 Fanari sedimentary unit lies on top of a thick cataclastic body (Figs. 3c and 3d) superimposed

312 over a 300-500 m-thick ductile strain gradient developed within both the Raches and

313 Karkinagrion granites (Laurent et al., 2015). Besides, the initial geological outline of the

314 Karkinagrion granite (Ring, 2007) was significantly modified and extended toward the north

315 (Fig. 2). Detailed field work within and around this granite massif allows the recognition of a

316 large-scale migmatite complex closely associated with this S-type granite (see recent

317 description in Laurent et al. (2015)) (Fig. 3e).

318

319 Accepted Manuscript

320 4. Structural analysis of the ductile deformation in metamorphic rocks

321

322 4.1. Main planar fabrics

323

Page 14 of 86 324 All metamorphic rocks and most of magmatic lithologies from Ikaria are pervasively

325 foliated. In most cases, compositional layering in the metasediments has been transposed into

326 a main, generally shallow-angle foliation.

327 221 foliation planes were measured in all lithologies of the two metamorphic units.

328 Besides, bedding was measured in Fanari unit. Measurements, statistical analysis and foliation

329 trajectories are reported on Figs. 4a, b and c, respectively. The dip of foliation planes displays

330 a large range of variation between 10° and 50° for a mean value at 20-25° (Fig. 4b).

331 Smoothing out these small-scale dip variations, the main foliation planes commonly dip away

332 from the long axis of the island. The strike of foliation therefore shows a fairly concentric

333 pattern depicting a NE-SW elongated structural dome (Fig. 4c), which ends in the northeast.

334 The dome axis, issued from the inversion of field measurements, trends N037°E, which

335 corresponds to the main orientation of regional foliation (Figs. 4b and 4c). Additionally, this

336 dome presents a marked asymmetry with a steeper, 30-40° dipping southeastern flank (Fig.

337 4b). The deeper parts of the dome crop out in the central part of the southern coast of the

338 island, close to the contact with the Raches granite. Toward the southwest, the axis extends

339 seaward and turns to a more E-W orientation to finally show up again across the main

340 occurrence of migmatites (Fig. 4c). Strike and dip of the main foliation are generally

341 discontinuous across intrusive contact of the Raches intrusion, particularly to the south.

342 Conversely, foliation trajectories are more continuous to the north where both the granite and

343 the wall-rocks locally present an ultramylonitic fabric (Fig. 4c).

344 Accepted Manuscript

345

346 4.2. Stretching lineation

347

Page 15 of 86 348 Stretching lineation is carried by the foliation plane in most metamorphic rocks, while

349 it may sometimes occur as the unique strain marker in the intrusive rocks. It is marked by

350 various indicators, depending on lithology, metamorphic grade and strain intensity. In calcite

351 and dolomite marbles, it is defined by very fine-grained mica slates and the elongation of

352 graphite-rich inclusions. In metapelites, it is generally defined by elongated quartz rods and

353 phyllosilicate aggregates. It is also sometimes marked by the elongation and the truncation of

354 epidote or tourmaline in granitic rocks. Evidence for stretching is also recorded by the

355 preferred elongation and the brittle truncation of clasts in conglomeratic layers at the base of

356 Fanari unit, just above the Fanari detachment.

357 202 stretching lineations have been measured in the field in all metamorphic rocks.

358 Data are all reported in Fig. 5 together with some first-order statistics. At first glance, the

359 trend of lineation shows very little dispersion. It is centered on an average value of N008°E

360 ranging from N160°E and N020°E (Fig. 5b). Wrapped around the dome, the lineation plunges

361 to the north in the northwestern flank of the island and to the south in the southeastern flank.

362 At large-scale, the trend of stretching lineation describes slightly curved patterns from N-S to

363 more NNE-SSW directions. The spatial rotation of the stretching lineation can be correlated

364 with the relative structural position; NNE-SSW orientations are observed in the uppermost

365 parts of the metamorphic succession and particularly in the vicinity of the Fanari shear zone,

366 in the northeast of Ikaria or at Gialiskari (Fig. 5a).

367

368 Accepted Manuscript

369 4.3. Asymmetry of ductile deformation

370

371 The whole volume of the metamorphic succession and a great part of the magmatic

372 intrusions of Ikaria are pervasively affected by a top-to-the-N to -NNE ductile deformation

Page 16 of 86 373 (Fig. 5a). Kinematics indicators of top-to-the-N deformation are very common and often

374 unambiguous, particularly toward the top of Ikaria unit and in the bulk of Agios-Kirykos unit.

375 Differing in terms of style, asymmetry and physical conditions of deformation, descriptions in

376 Ikaria and Agios-Kirykos units are presented separately.

377

378

379 4.3.1. Top-to-the-N shearing gradient in Ikaria unit

380

381 Shear bands accompanied by asymmetric boudinage are the most common kinematic

382 indicators in Ikaria unit (Fig. 6). Top-to-the-N shear bands are particularly abundant in

383 metapelite layers interleaved with more competent rocks such as marbles, metaquartzites or

384 metabasites. Due to asymmetric dome-shaped architecture of the island, shear bands display,

385 in their present position, gentle to moderate north-westward dips and normal-sense

386 displacements on the northwestern flank of the island while they often present flat or even

387 “reverse” geometry on the southeastern flank. Boudinage occurs at various scales in

388 alternating lithologies such as metapelites interleaved with marbles, metaquartzites or

389 metabasites. Boudins frequently show asymmetric shapes consistent with a northward

390 asymmetry. However, antithetic bookshelf structures compatible with a top-to-the-N sense of

391 shear can be developed within marble at different scales as observed by Ring (2007).

392 Top-to-the-N ductile deformation appears unevenly distributed within Ikaria unit,

393 primarily controlledAccepted by the relative structural position. Manuscript Evolution of its asymmetry can be

394 studied along a composite cross-section from the deepest parts to the top of this unit from

395 Plagia to . The deepest parts crop out on the southeastern coast between Chrisostomos

396 and Plagia. There, the most typical structural feature is an intense folding of the primary

397 compositional layering (Fig. 6a). Folds display a wide range of morphologies but all share

Page 17 of 86 398 common subhorizontal axial planes consistent with vertical flattening. A N-S stretching

399 lineation is only developed in metapelitic or sometimes quartzitic layers while other

400 lithologies display more randomly oriented intersection lineations. Sense of shear is often

401 ambiguous with the presence of both top-to-the-N and top-to-the-S shear criteria suggesting a

402 strong flattening component. Upward, the development of meter-scale decimeter-thick shear

403 bands marks the appearance of a clear asymmetry. Shear bands delimit asymmetric boudins

404 that preserve more ductile deformation as tight recumbent to isoclinal folds, or only as

405 detached fold hinges (Fig. 6b). However, at variance with deeper levels of the unit, fold axes

406 trend parallel to the stretching lineation arguing for a significant component of shearing in the

407 direction of stretching (Fig. 6c). Upward, along with the multiplication of shear bands, the

408 decreasing size of shear domains and boudins, metapelitic rocks become more homogeneous.

409 Lenses of metabasites, dolomitic marbles and remains of metaquartzites occur as sigma or

410 more rarely delta-type porphyroclasts systems consistent with an overall top-to-the-N sense of

411 shear (Fig. 6d). Intense shear strain has turned the rocks into fine-grained mylonites (Fig. 6e).

412 The resulting cross-section shows a first-order strain gradient, where asymmetric stretching is

413 more and more systematic toward the highest structural levels of the unit (Fig. 6f).

414 Metamorphic conditions that prevailed during deformation also depend on structural

415 position. Amphibolite-facies assemblages and gneisses are generally well preserved in the

416 deepest parts of Ikaria unit. There, first retrogression stages are clearly synkinematic, as

417 exemplified by the crystallization of biotite around stable garnets forming sigma-type

418 porphyroclasts systemAccepted (Fig. 7a). Upward, developed Manuscript during garnet breakdown, asymmetric

419 strain shadows around garnet contain chlorite and white-micas. Biotite, still metastable in the

420 bulk of the rock is retrograded to chlorite in shear bands. Associated with the crystallization

421 of large amounts of synkinematic chlorite and albite, top-to-the-N shearing in the highest

422 levels was clearly recorded within greenschist-facies metamorphic conditions (Fig. 7b).

Page 18 of 86 423

424

425 4.3.2. Deformation in Agios-Kirykos unit and along Fanari shear zone

426

427 The Agios-Kirykos unit can be studied along a series of foliation-orthogonal valleys

428 from Therma to Agios Kiriaki (see location on Fig. 2). It consists in the alternation of marble

429 and dark metapelites layers reaching 150-200 m of structural thickness. On most outcrops,

430 top-to-the-NNE ductile shearing is clear. Greenschist-facies shear bands are abundant and

431 display a clear upward evolution to more localized or even cataclastic flow (compare Fig. 8a

432 and Fig. 8b). The core of Fanari shear zone, developed at the expense of the uppermost parts

433 of Ikaria unit, is well exposed in the western part of Agios Kiriaki harbor (Fig. 8c). While the

434 contact itself is hidden by the airport runaway, Fanari unit crops out directly to the northeast,

435 100 m apart (Fig. 8c). There, rocks present a northeast-dipping mylonitic fabric and carry a

436 N035°E stretching lineation marked by numerous quartz rods and boudins of all sizes from a

437 few centimeters to several meters. Marble layers are stretched and boudinaged within the

438 metapelitic matrix. The internal deformation of marbles displays successive stages of

439 symmetric boudinage, evolving from ductile to brittle (Fig. 8d). Metapelite levels show a

440 clear non-coaxial component of shearing consistent with an overall top-to-the-NNE

441 kinematics (Fig. 8e). These layers, generally thinned to 1 m or even less, display a single and

442 penetrative set of shear bands locally obliterating the main foliation. Spacing between shear

443 bands, controlledAccepted by the presence of quartz lenses Manuscript is locally as small as 1-3 cm (Fig. 8e).

444 Chlorite is abundant within these rocks and adopts a clear synkinematic character in the

445 vicinity of shear bands. The brittle expression of the Fanari shear zone is not exposed there

446 but occurs spectacularly near Fanari, 1 km further south along the coast (Fig. 9). The Fanari

447 detachment is described in the next section.

Page 19 of 86 448

449

450 5. Brittle deformation analysis

451

452 Ikaria presents either large to meso-scale brittle structures or locally pervasive

453 networks of small-scale brittle faults systems developed from the ductile-brittle transition to

454 the brittle field. Brittle structures of all scales are first described. Results of the paleostress

455 analysis are then presented. It must be noted here that some NNE-SSW lineaments, visible on

456 satellite images, do not have any particular expression in the field.

457

458

459 5.1. Description of brittle structures

460

461 The Fanari detachment is the only large-scale structure that was active under both

462 ductile and subsequent brittle conditions. Displacement in the brittle regime over the Fanari

463 detachment is attested by the development of cataclasites particularly well exposed along the

464 southeastern coast of the island as a series of several hundreds of meters of continuous

465 outcrops (Fig. 9). There, the detachment plane appears stripped of its sedimentary cover over

466 a large surface revealing large-scale corrugations and crescent-shaped structures that remind

467 the Cape outcrop on Mykonos (Lecomte et al., 2010; Menant et al., 2013) or the Platy

468 Gialos outcrop onAccepted Serifos (Grasemann and Petrakakis, Manuscript 2007). Breccias, locally reaching more

469 than 2 m, are mainly developed at the expense of Agios-Kirykos unit but also from the very

470 first decimeters of Fanari unit. Cataclastic rocks often show typical clasts size ranging from 1

471 to 20 cm embedded in reddish-brownish cement supposed to be composed of Fe-rich oxy-

472 hydroxides and carbonates. The detachment plane displays a NE-SW strike and a 60°

Page 20 of 86 473 southeast-dip. It carries shallow-dipping, southwest-plunging, large-scale corrugations

474 consistent with the opening of perpendicular cracks. Variable types of kinematics indicators

475 of top-to-the-NNE brittle deformation in the detachment footwall unambiguously ascribe a

476 left-lateral reverse sense of displacement. A secondary discrete striae indicates locally a

477 reverse sense of displacement. At the scale of the map, kinematics of brittle motion over

478 Fanari detachment indicates northeast-directed displacement consistent with the kinematics of

479 the last increments of ductile deformation (compare kinematics given on Fig. 9 with the last

480 ductile stretching lineation on Fig. 5). In its current position, the detachment plane is offset by

481 a set of NW-SE vertical faults carrying oblique striations.

482 Cataclastic rocks also occur on the northern part of Ikaria at Gialiskari (Figs. 3c and

483 3d). They overprint the uppermost parts of a 500 m-thick strain gradient developed

484 exclusively at the expense of the underlying Raches granite (Laurent et al., 2015). Here, they

485 bound again a tectonic unit considered as the lateral equivalent of Fanari unit exposed further

486 east. Sediments, consisting on alternating grey sandstones, conglomerates and brownish

487 limestones are heavily affected by extensional features. WNW-ESE normal faults are

488 organized in conjugate sets accompanied by a dense array of subvertical WNW-ESE veins

489 that both overprint a gently south-dipping bedding.

490 Beside this major brittle structure, small-scale brittle features mostly correspond to late

491 W-E to NW-SE conjugate faults systems developed in all lithologies (Figs. 10 and 11).

492 Displacement on faults ranges typically from a few centimeters to a few meters and can

493 generally be constrainedAccepted in the field using marker -levels.Manuscript In the entire studied area, the fault

494 population is dominated by dihedral orientations arguing for the formation of newly formed

495 conjugate sets of faults rather than reactivated shear planes. Normal faults are often

496 accompanied by subvertical joints and tension gashes filled with quartz, chlorite, iron oxy-

497 hydroxides in the metamorphic rocks, quartz, chlorite or even epidote and tourmaline in

Page 21 of 86 498 granites and mostly calcite in the sediments. The close association between faults and joints

499 often displays contradictory intersection relationships and thus argues for a contemporaneous

500 development.

501

502

503 5.2. Inversion of fault-slip data

504

505 Paleostress orientation patterns were evaluated by the Win-Tensor computer-aided

506 inversion software (Delvaux and Sperner, 2003). The reduced paleostress tensor consists in

507 the identification of orientation of the three principal stress axes and the axial ratio of stress

508 ellipsoid. Determination of the paleostress axes was completed by the analysis of

509 accompanying brittle structures such as joints, tension gashes or stretched pebbles. Brittle

510 structure analysis was conducted over 9 main sites scattered over the island (see Fig. 5 for

511 location). Results are presented in the next two sections.

512

513

514 5.2.1. Using microstructures to unfold the Fanari detachment

515

516 Although the Fanari detachment consistently exhibits top-to-the-NE kinematics, it

517 currently crops out with the geometry of southeast-dipping left-lateral reverse fault zone, a

518 quite rare configuratiAcceptedon in the Aegean domain (Fig.Manuscript 9). A detailed study of geometrical

519 relationships between detachment, sediments of Fanari unit and attitude of small-scale brittle

520 structures affecting either the sediments or the detachment plane, allows proposing a

521 restoration scenario of the initial geometry of the Fanari detachment (Fig. 10a). Two reference

522 outcrops (6 and 7) were selected 2 km apart within sediments of Fanari unit along the

Page 22 of 86 523 detachment (see position on Fig. 5) with contrasting spatial relationships between structures.

524 Outcrop of station 7 (Fig. 10b), which presents almost vertical bedding, displays a conjugate

525 set of subvertical faults affecting alternating sandstones and conglomeratic layers. N030°E to

526 N160°E faults present consistent sinistral kinematics while N070°E to N130°E faults present

527 dextral kinematics. Part of these faults even cuts across the Fanari detachment causing 1-5 m

528 offsets of the plane. Outcrop of station 6 also presents two sets of faults that affect a 40°

529 dipping bedding further southeastward. The dominant set of faults displays subvertical fault

530 planes carrying both normal and reverse kinematics. This set is accompanied by a subordinate

531 set of flat to gently-dipping fault planes that also present both normal and reverse kinematics

532 (Fig. 10b). In their present geometry, this heterogeneous fault set requires the superimposition

533 of two distinct stress regimes. Back-tilting of those systems of faults about a horizontal axis,

534 in order to obtain a horizontal bedding, permits to get a coherent stress regimes for both sites

535 6 and 7. Maximum principal stress axis becomes vertical and the two others become

536 horizontal, compatibly with a NE-SW extension for both sites. It strongly suggests that faults

537 form before an unequal tilting of the different outcrops. This assumption is confirmed further

538 to the southwest where both the detachment plane and the bedding of sediments are

539 subhorizontal (i.e. less than 20° toward the southwest) while tension-gashes are now

540 subvertical and all faults appear as a single conjugate normal fault set preserved in its initial

541 attitude. Importantly, back-tilting of the whole system, including the Fanari detachment,

542 results in a system where the detachment plane operates at shallow angle, 10-15° to the

543 northwest, with Accepted normal-dextral-sense kinematics andManuscript cuts the sediments down-section as

544 observed in the neighboring island of Mykonos (e.g. Lecomte et al., 2010; Menant et al.,

545 2013). In this restored position (Fig. 10a), sediments are pervasively affected by normal

546 faulting consistent with a unique and common NNE-SSW to NE-SW extension (Fig. 10b).

547 This paleostress solution is consistent with the paleostress field deduced from other stations

Page 23 of 86 548 throughout the island and previous studies (e.g. Kumerics et al., 2005). This point is discussed

549 below. The causes of the Fanari detachment tilting and the more general arching of the ductile

550 to brittle fabrics at the scale of the whole island are addressed in the discussion section.

551

552

553 5.2.2. Large-scale consistency of the paleostress inversions

554

555 Despite the local coexistence of both N80-110°E and N130-160°E trending normal

556 faults, results present rather small internal dispersion (Fig. 11). At most stations, stress tensor

557 analysis shows a consistent subvertical orientation for the maximum principal stress axis (1).

558 In other stations (e.g. sites 6 and 7 on Fig. 10b), 1 was restored using horizontal-axis back-

559 tilting rotation. The minimum principal stress axis (3) is horizontal or very gently dipping

560 with a consistent NNE-SSW to more NE-SW direction of extension. In turn, the overall

561 consistency of the direction of 3 prevents from significant vertical-axis large-scale block

562 rotations. Planes of joint and tension gashes correspond to the calculated 1-2 plane, and

563 their poles therefore appear scattered around the 3 axis. Brittle structures recorded during the

564 last exhumation stages present a marked consistency of a NE-SW stretching direction

565 occurring in an extensional regime.

566 567 Accepted Manuscript 568 6. Thermal structure

569

570 The RSCM method is based on the quantitative study of the degree of organization of

571 carbonaceous material (CM), which is a reliable indicator of metamorphic temperature (e.g.

572 Pasteris and Wopenka, 1991; Beyssac et al., 2002; Lahfid et al., 2010). Because of the

Page 24 of 86 573 irreversible character of graphitization, CM structure is not sensitive to the retrograde

574 reactions and permits to determine peak temperature conditions (Tmax) reached during

575 metamorphism, using an area ratio (R2 ratio) of different peak of Raman spectra (Beyssac et

576 al., 2002). Tmax can be determined in the 300-640 °C range with an accuracy of ± 50 °C

577 related with the precision and the dispersion of petrological data used for the method

578 calibration (Beyssac et al., 2002). Relative uncertainties on Tmax are however much smaller,

579 around 10-15 °C and relative variations of that order of magnitude can be detected (e.g.

580 Gabalda et al., 2009; Vitale Brovarone et al., 2013; Augier et al., 2015). Sampling was

581 performed in order to quantitatively describe the large-scale thermal structure of Ikaria to

582 complement the few existing, punctual P-T estimates (Kumerics et al., 2005; Martin et al.,

583 2011). Sampling was therefore regularly distributed within the tectonometamorphic

584 succession of Ikaria and a more systematic sampling was carried out in the vicinity of possible

585 second-order thermal effects such as intrusions and tectonic contacts. 35 samples consisting of

586 CM-bearing metasediments were collected (see Fig. 12 for location). In order to bring out and

587 possibly smooth out the inner structural heterogeneity of CM within samples, 13 to 20 spectra

588 were recorded for each sample. Detailed results, including R2 ratio, number of spectra, Tmax

589 and standard deviation are presented on Table 1. In addition, RSCM temperatures are all

590 reported on Fig. 12.

591 RSCM results embrace a wide range of temperature from 391 to 625 °C (Table 1). At

592 first glance, Tmax presents a correlation with the relative structural position. 625 °C was 593 indeed recorded Acceptedfor the deepest parts of the structural Manuscript dome with gneisses while 390 °C was

594 retrieved for the uppermost parts of the metamorphic succession where low-crystallinity

595 micaschists crop out. Ikaria therefore appears as a metamorphic dome in which temperature

596 increases down-section (Figs. 12a and 12b). Samples from Ikaria unit show rather high Tmax,

597 from 625 to 500 °C, presenting an important scatter. The 550 °C temperatures for the volume

Page 25 of 86 598 of Ikaria unit are moreover consistent with the amphibolite-facies metamorphic associations.

599 Samples from Agios-Kirykos unit display lower temperature ranging between 450 and 390 °C

600 (Table 1). Two main sets of temperatures can then be identified separated by a 50 °C

601 temperature gap (Figs. 12a and 12b). This gap of temperature corresponds to the position of

602 the Agios-Kirykos shear zone, independently recognized on the basis of structural criteria.

603 This particular feature was studied in the east of Ikaria, and particularly along a foliation-

604 orthogonal cross-section upstream of Therma valley (Fig. 12b). There, Tmax shows a regular

605 temperature decrease from 600 to 400 °C accompanied by a 50 °C normal-sense metamorphic

606 gap. Among the causes of temperature variation in Ikaria unit, local heating by intrusive rocks

607 has been tested in the vicinity of the Raches granite (Fig. 12c). Samples were picked at

608 decreasing distance from the intrusive contact for almost the same relative structural position.

609 Results show regularly decreasing temperatures from 580 °C to 520 °C, in accordance with

610 the intrusive character of the Raches granite (Fig. 3a). Accordingly, smaller intrusive bodies

611 as the Kefala diorite or small-scale pegmatite dykes that are only partly mapped (Fig. 2) may

612 be responsible of erratic, isolated high Tmax results.

613

614

615 7. Age constrains on the partial melting event

616

617 7.1. Analytical methods

618 Accepted Manuscript

619 Analyses were performed in thin section by laser ablation inductively coupled plasma

620 spectrometry (LA-ICPMS) at the Laboratoire Magmas et Volcans (LMV), Clermont-Ferrand

621 (France). The ablation is performed using a Resonetics Resolution M-50E system equipped

622 with an ultra-short pulse (<4ns) ATL excimer 193 nm wavelength laser. This laser system is

Page 26 of 86 623 coupled with Agilent 7500 cs ICP-MS equipped with a pumping system to enhance the

624 sensitivity. Spot diameter of 9 µm was used with a 1 Hz repetition rates. Ablated material is

625 transported using a helium flux, and then mixed with nitrogen and argon before being injected

626 into the plasma source. Analytical procedures for monazite dating are reported in details in

627 Didier et al. (2013), Didier et al. (2014) and Paquette and Tiepolo (2007). Following isotopes

628 204(Pb+Hg), 206Pb, 207Pb, 208Pb, 232Th and 238U were acquired. Data disturbed by inclusions,

629 fractures or age mixing (between different areas of a single grain) were not taken into account

630 for calculation. The occurrence of initial Pb in the sample can be monitored by the evolution

631 of the 204(Pb+Hg) signal intensity, but no Pb correction was applied. Elemental fractionation

632 and mass bias were corrected making repeated analyses on Trebilcock monazite (272 ±2 Ma,

633 Tomascak et al., 1996). Analyses on the Moacyr monazite (Cruz et al. 1996; Seydoux-

634 Guillaume et al. 2002; Gasquet et al. 2010; Fletcher et al. 2010) at the beginning and at the

635 end of each session, treated as unknowns, verify the reproducibility, especially for the

636 208Pb/232Th ages, and the accuracy of the corrections. Data reduction was carried out with

637 GLITTER® software package (van Achterbergh et al. 2001; Jackson et al. 2004). Calculated

638 ratios were exported and age diagrams were generated using Isoplot software package by

639 Ludwig (2001).

640 In this study, 208Pb/232Th ages are preferably used because: 1) U decay series could be

641 in disequilibrium in young monazites (Schärer, 1984), resulting in overestimated 206Pb/238U

642 ages; 2) 232Th is so abundant that 208Pb originating from initial Pb is negligible compared to

643 radiogenic 208PbAccepted (Janots et al. 2012; Didier et al. 2013).Manuscript However, in this study, 206Pb/238U

644 ages are fully consistent with the 208Pb/232Th ages suggesting the absence of disequilibrium in

645 the U decay series. Slight common Pb contamination is suggested by the Tera Wasserburg

646 diagram, enhanced by the large uncertainty of the 207Pb/235U ages due to very low 207Pb

647 content in young monazite ages.

Page 27 of 86 648

649

650 7.2. Results

651

652 Fourteen monazites have been analyzed for U-Th-Pb in situ dating in thin section from

653 one migmatite sample (IK01A; see Fig. 2 for localization; Qz, Pl, Afs, Bt, Sil and Zrn).

654 Monazites often appear as inclusions in biotite or set at grain boundaries and occur as large

655 pristine crystals (Fig. 13a; between 60 and 100 µm) showing in few cases irregular

656 boundaries. Besides, most of the crystals present a clear core and rim texture (Fig. 13b) which

657 represents variable Y and U contents : the core is Y and U-poor (Y2O3 between 0.3 and 1.6

658 wt% and UO2 between 0.2 and 0.8 wt%) whereas the rim is Y and U-rich (Y2O3 between 1.7

659 and 4.2 wt% and UO2 up to 3.7 wt %).

660 Thirty two analyses have been performed in the different domains of the monazite

661 grains. 208Pb/232Th analyses yielded ages between 14.9 ± 0.5 and 16.4 ± 0.5 Ma, with a

662 weighted mean age at 15.7 ± 0.2 Ma (MSWD = 4,1; N=32) . All ages appear always

663 concordant with the 206Pb/238U ages (Fig. 13c). The U-Pb ages are mostly discordant and

664 define a linear trend crosscutting the Concordia at 15.1 ± 0.2 Ma in a Tera Wasserburg

665 diagram (Fig. 13d). This suggests a slight common Pb contamination and the large uncertainty

666 of the 207Pb/235U ages due to the very low 207Pb content in young monazite. For this reason,

667 only 208Pb/232Th ages are considered here (see details in analytical methods).

668 The scatteringAccepted of the 208Pb/232Th ages (high ManuscriptMSWD) is broadly correlated to the core

669 to rim zonation, the Y and U-poor cores yielding older ages than the Y and U-rich rims (Fig.

670 13b). However, the small difference between the ages measured in the cores and in the rim of

671 the monazite grains (less than 1.5 Ma) does not enable a clear distinction between two age

672 groups. Because i) the monazite is well known to be an efficient chronometer in dating the

Page 28 of 86 673 high temperature metamorphic processes, and ii) no inheritance (suggesting the recording of

674 older metamorphic event) has been observed, these results clearly establish a Langhian age

675 for the partial melting event that affected the infrastructure of Ikaria Island.

676

677

678 8. Discussion

679

680 8.1. An overall top-to-the-N sense of shear as the main deformation record

681

682 The whole tectonometamorphic succession of Ikaria is pervasively affected by top-to-

683 the-N to -NNE ductile deformation whose intensity increases toward the top of the structure

684 (Figs. 6f and 14a). While the deepest parts of the tectonometamorphic succession exhibits

685 rather symmetrical deformation dominated by flattening, top-to-the-N shearing deformation

686 shows a large-scale strain gradient toward the Fanari shear zone along which most of the

687 strain is concentrated, with an upward evolution toward a more top-to-the-NNE kinematics

688 (Fig. 5). Characterized by a subordinate top-to-the-NNE strain gradient, the Agios-Kirykos

689 shear zone is furthermore responsible for a 50 °C gap of normal sense, as retrieved by RSCM

690 analyses (Fig. 12). These embedded gradients are accompanied by greenschist-facies

691 retrogression gradients quite obvious at the scale of the outcrop or hand-specimen (Fig. 7).

692 Crossing the ductile to brittle transition, the Fanari shear zone has even recorded increments

693 of motion in brittleAccepted conditions responsible for the Manuscript development of thick cataclasite bodies

694 (Figs. 3 and 9). The direction of ductile stretching and markers of later brittle extensional

695 motions on the detachment planes show a continuum of NNE-SSW stretching and top-to-the-

696 NNE shearing, consistent, in a broad sense, with extensional direction in both metamorphic

697 and sedimentary rocks (Fig. 11). This evolution is interpreted as an evidence of continuous

Page 29 of 86 698 stretching from middle to upper crustal levels through the ductile-brittle transition and

699 continuous shearing along a major detachment localizing deformation.

700 Low-temperature geochronology pointed out a clear northward younging of ZFT ages

701 (10.3 ± 0.3 to 7.1 ± 0.3 Ma) and AFT ages (8.4 ± 0.8 to 5.9 ± 0.8 Ma) consistent with an

702 overall top-to-the-N unroofing with apparent slip-rates over the Fanari detachment of 10-8

703 km/Ma (Fig. 14b; Kumerics et al., 2005). U-Th/He analyses on apatite for both flanks of the

704 dome yielded similar 6 Ma ages. Motion over the Fanari detachment seems stop near 6 Ma.

705 These ages constrain, in turn, the maximum age for the onset of passive “folding” of the

706 Fanari detachment plane around the dome, which postdates the last brittle motions of the

707 detachment and even the brittle normal faulting that affects sediments.

708 A progressive rotation of the stretching and shearing direction is observed from the

709 core to the outer flanks of the domes (Fig. 5). As suggested by the structural study, the

710 shearing strain has progressively localized through time upward from a distributed flow,

711 recorded in the amphibolite-facies conditions, to mylonites, ultramylonites and then

712 cataclasites in the vicinity of Fanari detachment (Figs. 3 and 9). Accordingly, the N-S

713 trending lineation preserved in the core is older than the NNE-SSW trending one at the top

714 along the main shear zones. The core of the dome has therefore rotated 35° counterclockwise

715 during its exhumation before the final doming that seems to be the latest event recorded on

716 Ikaria. This late event tends to confer to the dome the geometry of an a-type dome where the

717 stretching direction is parallel to the dome axis (Jolivet et al., 2004a) and which can suggests

718 an oblique compAcceptedonent of shearing during doming Manuscript (Le Pourhiet et al., 2012). A significant

719 component of E-W shortening is currently recorded in the deformation of the northern Aegean

720 domain as a result of westward motion of Anatolia along the North Anatolian Fault (NAF)

721 (e.g. Le Pichon and Kreemer, 2010). If the NAF has reached the northern Aegean some 6 Ma

722 ago as a localized crustal-scale fault zone (Armijo et al., 1999; Melinte-Dobrinescu et al.,

Page 30 of 86 723 2009), it has been suggested that dextral movements has been active since 12 Ma (Sengör et

724 al., 2005). In the field, consequences of an E-W shortening are clearly recorded in the

725 northern Cyclades, all the way to western Turkey (Angelier, 1976; Buick, 1991; Bozkurt and

726 Park, 1997; Ring et al., 1999; Avigad et al., 2001; Menant et al., 2013). This context, showing

727 a combination of exhumation, strike-slip and contemporaneous E-W shortening and N-S

728 crustal stretching, suggests a component of transtension during the last evolution stages (Le

729 Pourhiet et al., 2012; Fossen et al., 2013). This oblique component fits the hypothesis of a

730 left-lateral transfer zone from the Menderes massif to the Cyclades proposed by Ring et al.

731 (1999) and recently reassessed by Gessner et al. (2013). Detailed studies on the timing of the

732 formation of stretching directions across the dome coupled with the timing of doming are

733 necessary to go any further.

734

735

736 8.2. Eastern extension of the North Cycladic Detachment System

737

738 Obvious structural and metamorphic similarities of the series of detachments from

739 Andros to Mykonos, with coeval top-to-the-N to -NNE kinematics, led Jolivet et al. (2010) to

740 mechanically merge them in a single crustal-scale extensional structure, called NCDS. A

741 probable correlation of the Fanari detachment of Kumerics et al. (2005) with the NCDS was

742 even proposed (Jolivet et al., 2010). The existence on Ikaria of two detachments, one purely

743 ductile and the Accepted other evolving from ductile to brittleManuscript conditions and the presence of the

744 synkinematic granite piercing through the dome are indeed reminiscent of the anatomy of the

745 NCDS as described on Andros, Tinos and Mykonos (Jolivet et al., 2010). On Tinos (e.g.

746 Jolivet and Patriat, 1999; Jolivet et al., 2004a; Brichau et al., 2007) as well as on Mykonos

747 (e.g. Faure et al., 1991; Lecomte et al., 2010; Denèle et al., 2011; Menant et al., 2013),

Page 31 of 86 748 increments of ductile and then brittle deformation are recorded by synkinematic intrusions

749 emplaced into the main extensional shear zones. The Tinos Detachment is mostly ductile and

750 is pierced by the Tinos intrusion. Conversely, the Livada Detachment shows ductile then

751 brittle deformation and affects the Tinos intrusion (Brichau et al., 2007). This last detachment

752 is then pierced by the Mykonos intrusion. The more superficial Mykonos Detachment that

753 clearly post-dates all intrusions, operated only in purely brittle conditions and carries syn-

754 extension sediments (e.g. Lecomte et al., 2010; Menant et al., 2013). These three detachments

755 that operate sequentially form the NCDS. A similar evolution is observed on Ikaria.

756 Exclusively operating in the ductile field, the Agios-Kirykos shear zone is later cut down-

757 section by the Fanari detachment that evolves from ductile to brittle and even controls the

758 deposition of syn-tectonic sediments in the hangingwall.

759 Along with the current lack of precise time-constraints on synkinematic minerals,

760 timing of extensional deformation onset is still unclear on Ikaria. If partial melting occurred at

761 ca. 15.5 Ma, as showed in this study, the signification of the 25-17 K/Ar ages on hornblende

762 (Altherr et al., 1982) are questioned. These ages may reflect either Ar inheritance from an

763 earlier tectonometamorphic event or simply excess argon (i.e. extraneous argon), as suggested

764 by the strong scattering of these ages. Then, onset of extension might starts at or just after ca.

765 15.5 Ma. The 12-9 Ma K/Ar and Ar/Ar ages obtained on fabric-forming white-micas and

766 biotite are believed to record either a fast cooling or last recrystallizations during

767 mylonitization (see compilation on Fig. 14b; Altherr et al., 1982; Kumerics et al., 2005). In

768 parallel, the threeAccepted intrusions (Raches, Karkinagrion Manuscriptand Xylosyrtis) that crystallized at 14-13

769 Ma therefore synkinematically emplace within the Fanari shear zone (U-Pb on zircon; Bolhar

770 et al., 2010; Laurent et al., 2015). Intrusions all underwent a common fast cooling to

771 temperatures close to the ductile to brittle transition (Stöckhert et al., 1999; Imber et al., 2001)

772 at ca. 10 Ma (K/Ar and Ar/Ar on micas; Altherr et al., 1982; Kumerics et al., 2005). Being

Page 32 of 86 773 only active in the ductile field, the Agios-Kirykos therefore probably ceased to be active some

774 10 Ma ago. Displacement and further strain localization were thus transferred to the Fanari

775 shear zone, responsible for the final exhumation, in the ductile and then the brittle field. The

776 last exhumation stages were characterized by fast (100-80 °C/Ma) cooling rates and ended at

777 around 6 Ma (Fig. 14b; FT and U-Th/He on zircon and apatite; Kumerics et al., 2005). As a

778 comparison, crystallization and cooling of the Mykonos pluton occurred between 13.5 Ma and

779 9 Ma as a result of the fast exhumation related to the activity of the Livada and Mykonos

780 detachments (e.g. Brichau et al., 2008; Lecomte et al., 2010; Jolivet et al., 2010). Low-

781 temperature ages even suggest that motions over the Fanari detachment continued until 6 Ma

782 and are thus the last top-to-the-N to -NNE detachment active in the Aegean domain. These

783 similarities in terms of geometries, kinematics and timing of the Ikaria detachments and the

784 various branches of the NCDS confirm the proposed eastward extension of the NCDS all the

785 way to the eastern end of Ikaria in the Agios-Kirykos and Fanari shear zones (Fig. 15).

786 Similarly to Mykonos, the detachment system of Ikaria, does not exhume HP-LT

787 metamorphic rocks belonging to the Cycladic Blueschists unit, at variance with Tinos and

788 Andros. Indeed, along with the structural position and the lithostratigraphic succession, the

789 apparent lack of any HP-LT imprint either in Ikaria and Agios-Kirykos units precludes

790 correlations with the Cycladic Blueschists unit. The Ikaria and Agios-Kirykos units are

791 characterized by the same monotonous lithologies and a similar 80-100 °C/km thermal field

792 gradient, it is proposed that both units were equilibrated along a single, warm gradient rooting

793 in partially-moltenAccepted rocks. Manuscript

794 Correlated with the Upper Cycladic unit, in a broad sense, the Fanari unit reworks

795 large amount of Pelagonian detritus shed into the extensional basins (Photiades, 2002a).

796 While deposition of sediments on the Upper Cycladic unit started around 23 Ma throughout

797 the Aegean domain (Angelier et al., 1976; Sánchez-Gómez et al., 2002; Kuhlemann et al.,

Page 33 of 86 798 2004), sediments preserved on Ikaria are attributed to Late Oligocene to Early Pliocene

799 (Papanikolaou, 1978; Photiades, 2002a; 2002b). This rather recent deposition age is consistent

800 with last denudation ages ascribed to last extensional motions over the Fanari detachment

801 (Kumerics et al., 2005), which may be responsible for maintaining a significant tectonic

802 subsidence.

803

804

805 8.3. Definition of the Ikaria MCC and the central Aegean HT zone

806

807 The structure of Ikaria is composed of two main parts separated by a detachment

808 system consisting in shallow-dipping extensional ductile to brittle shear zones connecting

809 surficial to mid-crustal levels (e.g. Fig. 2). The bulk finite architecture of the footwall consists

810 in an elongated structural dome roofed by a partly eroded mylonite, ultramylonite and

811 cataclasite carapace (Fig. 4). This geometrical dome is paired with a fairly concentric

812 distribution metamorphic zones of decreasing grade from migmatites and 625 °C in the core

813 to 390 °C in the externals parts (Fig. 12). The dome flanks display 500-m thick strain gradient

814 evolving upward to mylonites characterized by the uniformity of shear sense from one limb to

815 the other. They are partially overprinted by cataclasites and fault-rocks in the Fanari

816 detachment zone (Figs. 3 and 9). The detachment itself is marked by a clear-cut fault plane

817 carrying evidences for large-scale displacements in the brittle field (Fig. 9). The hangingwall

818 unit, exclusivelyAccepted composed sediments, is only characterized Manuscript by brittle deformation (Fig. 10).

819 Following this description, Ikaria dome shares all attributes of a typical migmatite-cored

820 MCC (see Platt et al., 2014, for review) firstly described in the Basin and Range of western

821 United States (Davis and Coney, 1979; Crittenden et al., 1980; Wernicke, 1981, 1985), and

822 then in other Alpine (e.g. Lister et al., 1984; Dewey, 1988; Gautier et al., 1993; Gautier and

Page 34 of 86 823 Brun, 1994; Jolivet and Patriat, 1999; Vanderhaeghe, 2004) and Variscan or Caledonian

824 orogenic belts (e.g. Norton, 1986; Andersen et al., 1991).

825 Partial melting is a key factor that controls the strength of the continental crust. The

826 recent discovery of migmatites in the infrastructure of the Ikaria MCC dome may have

827 important implications for the behavior of the Hellenic continental crust during Oligo-

828 Miocene extension. Several high-grade gneiss and migmatite massifs have already been

829 described in the Aegean realm. However, most of these occurrences of HT rocks are basement

830 rocks inherited from ancient tectonometamorphic events. An early Paleozoic HT event is

831 known in the Menderes (e.g. Gessner et al., 2004) and in the East of the Aegean Sea,

832 particularly on Ikaria (Kumerics et al., 2005; Ring et al., 2007). There, a ca. 460 Ma age

833 retrieved form a pegmatite dyke may reflect a minimum age for HT metamorphism (Kumerics

834 et al., 2005). A Variscan inheritance is also well described in the Cycladic basement unit in

835 the Southern Cyclades (e.g. Henjes-Kunst and Kreuzer, 1982; Andriessen et al., 1987; Keay

836 and Lister, 2002). On these islands, a ca. 320 Ma granite and amphibolite facies gneisses are

837 partially overprinted by the Eocene HP imprint (e.g. Henjes-Kunst and Kreuzer, 1982;

838 Andriessen et al., 1987; Keay and Lister, 2002; Huet et al., 2009; Augier et al., 2015). The

839 only Oligo-Miocene migmatites crop out in the core of the Naxos MCC, the first that has been

840 described in the Aegean domain (e. g. Lister et al., 1984). There, the age of the partial melting

841 is quite constrained by U-Pb analyses on zircon yielding series of ages ranging from 21 to 17

842 Ma (Keay et al., 2001). The age of migmatites recently discovered on Ikaria is therefore of

843 prime importance.Accepted In this study, pioneer U-Th-Pb analysesManuscript on monazites from a leucosome of

844 the migmatites yielded an age of 15.7 ± 0.2 Ma (Fig. 13). This result falls in the same age-

845 range than the recent U-Pb age on zircon ascribed to the Karkinagrion S-type granite (Bolhar

846 et al., 2010) together with closely associated migmatites (Laurent et al., 2015). Clearly

847 associated with the same large scale partial melting event that affected the Hellenic

Page 35 of 86 848 continental crust, anatexis at the latitude of Ikaria appears slightly younger than further south

849 on Naxos. Along with Naxos, Mykonos and Ikaria, HT MCCs all cluster in the central part of

850 the Aegean domain. It is here proposed to define the central Aegean HT zone as a large-scale

851 extensional tectonic window of high-grade, partially-molten rocks (Fig. 15), of Early-Middle

852 Miocene age (Keay et al., 2001; this study) (Fig. 15). There, intense stretching leads to the

853 complete tectonic omission of the Cycladic Blueschists unit that is observed in colder MCCs

854 further west in , Andros or Tinos (e.g. Trotet et al., 2001a; Mehl et al., 2007), to the east

855 in Samos (e.g. Ring et al., 1999) or to the south in Ios, Folegandros and Sikinos (e.g. Huet et

856 al., 2009; Augier et al., 2015). This gradient of finite stretching toward the center and east of

857 the Cyclades is attested by the joint evolution of the topography, the crustal thickness and the

858 first order transition from brittle to ductile deformation from marginal areas to the center of

859 the Aegean domain (e.g. Jolivet et al., 2010). Uprise of partially-molten materials in-between

860 less-extended Cycladic Blueschists unit evokes crustal-scale equivalents of the scar folds that

861 result from ‘‘boudins’’ separation in the process of boudinage (Tirel et al., 2008). This simple

862 model should however be adapted as three distinctive detachment bounding smaller-scale

863 MCCs are currently observed (Fig. 15). The NCDS (Jolivet et al., 2010) exhumes Andros,

864 Tinos, Mykonos and Ikaria in the north while the NDP exhumes Naxos and Paros in the

865 center (Buick, 1991; Gautier et al., 1993; Vanderhaeghe, 2004; Duchêne et al., 2006;

866 Kruckenberg et al., 2011). Both are associated with top-to-the-N shear. Finally, further

867 southwest, the alignment of Kea, Kythnos and Serifos shows another row of small-scale

868 domes exhumedAccepted by the top-to-the-SW WCDS (GrasemannManuscript et al., 2012). If top-to-the-S

869 criteria on Ios and Sikinos are related to a thrust (Huet et al., 2009; Augier et al., 2015) and

870 not to a detachment (Vandenberg and Lister, 1996; Forster and Lister, 1999; Thomson et al.,

871 2009; Ring et al., 2011) then extensional top-to-the-N shearing is systematically observed

872 from north to south of the central Aegean (Fig. 15). This result contrasts with the more

Page 36 of 86 873 symmetric deformation recognized in marginal areas, where symmetrically arranged

874 detachment systems depict more bivergent extensional systems either to the west (Grasemann

875 et al., 2012) or east in Samos (Ring et al., 1999) or in the Menderes massif (Hetzel et al.,

876 1995b; Gessner et al., 2001). A correlation between the amount of stretching of the crust and

877 the degree of non-coaxiality is therefore proposed. Just as the Tyrrhenian or the Alboran back-

878 arc systems, the causes of this regional-scale non-coaxial extension in the Aegean back-arc

879 domain is a first-order question that is outside the focus of this paper but may pertain to the

880 interactions between the upper plate, the retreating subducting slab and the flowing

881 asthenospheric mantle (Jolivet et al., 2009; Sternai et al., 2014).

882

883

884 9. Conclusions

885

886 Although Ikaria is one of the largest Aegean islands and paves the way between the

887 Cyclades and the Menderes massif, its geology and tectonic evolution remain poorly

888 understood since recent studies lead to very conflicting maps and interpretations. This paper

889 reports a detailed 3D study of the geometry and the thermal structure of Ikaria. The structural

890 study shows that the HT-LP foliation is arched, forming a NE-SW trending structural dome

891 cored by partially molten rocks and intruded by late intrusive granitic bodies. Lineation shows

892 a N-S to NNE-SSW ductile stretching associated with an overall top-to-the-N to -NNE sense

893 of shear. Final exhumationAccepted of the dome was completed Manuscript by Miocene extensional system made

894 of two main top-to-the-N to -NNE shear zones, operating in the ductile and then the brittle

895 fields. The thermal structure revealed by the RSCM approach strengthened the subdivision of

896 the metamorphic succession in two main metamorphic units (Ikaria and Agios-Kirykos units)

897 with an upward decrease of maximum temperature, separated by the Agios-Kirykos shear

Page 37 of 86 898 zone. The Fanari detachment permits to juxtapose the sedimentary Fanari unit directly on

899 metamorphic rocks. The distribution of RSCM temperatures within the dome and the presence

900 of migmatites of ca. 15.5 Ma age in the western part of the island fit the description of a HT

901 MCC such as Naxos or Mykonos. The proposed tectono-metamorphic evolution of the dome

902 is consistent with the evolution of the northern Aegean area controlled by the eastern

903 extension of the NCDS. Definition of Ikaria as a HT MCC allows the definition of the

904 geographic extent of the central Aegean HT zone associated with strictly asymmetric top-to-

905 the-N ductile shearing and an Early-Middle Miocene partial melting event. A correlation

906 between the amount of stretching of the crust and the degree of non-coaxiality is therefore

907 proposed.

908

909

910 Acknowledgements

911

912 Uwe Ring and Berhnard Grasemann are thanked for their constructive comments and

913 suggestions that improved the manuscript. This work has received funding from the European

914 Research Council (ERC) under the Seventh Framework Programme of the European Union

915 (ERC Advanced Grant, grant agreement No 290864, RHEOLITH) as well as Institut

916 Universitaire de France. It is a contribution of the Labex VOLTAIRE. The authors are

917 grateful to S. Janiec and J.G. Badin (ISTO) for the preparation of thin sections.

918 Accepted Manuscript

919

920 References

921

Page 38 of 86 922 Altherr, R., Kreuzer, H., Lenz, H., Wendt, I., Harre, W. & Dürr, S. (1994). Further Evidence

923 for a Late Cretaceous Low-pressure high-temperature Terrane in the Cyclades, Greece -

924 Petrology and Geochronology of Crystalline Rocks from the Islands of Donoussa and Ikaria.

925 Chemie Der Erde 54, 319–328.

926

927 Altherr, R., Kreuzer, H., Wendt, I., Lenz, H., Wagner, G. A., Keller, J., Harre, W. &

928 Höhndorf, A. (1982). A Late Oligocene/Early Miocene High Temperature Belt in the Attic-

929 Cycladic Crystalline Complex (SE Pelagonian, Greece). Geologisches Jahrbuch 23, 97–164.

930

931 Altherr, R., Schliestedt, M., Okrusch, M., Seidel, E., Kreuzer, H., Harre, W., Lenz, H., Wendt,

932 I. & Wagner, G. A. (1979). Geochronology of High-Pressure Rocks on Sifnos (cyclades,

933 Greece). Contributions to Mineralogy and Petrology 70, 245–255.

934

935 Altherr, R. & Siebel, W. (2002). I-type plutonism in a continental back-arc setting: Miocene

936 granitoids and monzonites from the central Aegean Sea, Greece. Contributions to Mineralogy

937 and Petrology 143, 397–415.

938

939 Andersen, T. B., Jamtveit, B., Dewey, J. F. & Swensson, E. (1991). Subduction and eduction

940 of continental crust: major mechanisms during continent-continent collision and orogenic

941 extensional collapse, a model based on the south Norwegian Caledonides. Terra Nova 3, 303–

942 310. Accepted Manuscript

943

944 Andriessen, P. A. M., Banga, G. & Hebeda, E. H. (1987). Isotopic age study of pre-Alpine

945 rocks in the basal units on Naxos, Sikinos and Ios, Greek Cyclades. Geologie en Mijnbouw

946 66, 3–14.

Page 39 of 86 947

948 Angelier, J. (1976). Sur l’alternance mio-plio-quaternaire de mouvements extensifs et

949 compressifs en Egée orientale: l’île de Samos (Grèce). Comptes Rendus de l’Academie des

950 Sciences de Paris 283, 463–466.

951

952 Armijo, R., Meyer, B., Hubert, A. & Barka, A. (1999). Westward propagation of the North

953 Anatolian fault into the northern Aegean: Timing and kinematics. Geology 27, 267–270.

954

955 Aubouin, J. & Dercourt, J. (1965). Sur la géologie de l’Egée: regard sur la Crète. Bulletin de

956 la Societe Geologique de France 7, 787–821.

957

958 Augier, R., Jolivet, L., Gadenne, L., Lahfid, A. & Driussi, O. (2015). Exhumation kinematics

959 of the Cycladic Blueschists unit and back-arc extension, insight from the Southern Cyclades

960 (Sikinos and Folegandros Islands, Greece). Tectonics 34, 152–185.

961

962 Avigad, D. & Garfunkel, Z. (1989). Low-angle faults above and below a blueschist belt:

963 Tinos Island, Cyclades, Greece. Terra Nova 1, 182–187.

964

965 Avigad, D. & Garfunkel, Z. (1991). Uplift and exhumation of high-pressure metamorphic

966 terrains: the example of the Cycladic blueschist belt (Aegean Sea). Tectonophysics 188, 357–

967 372. Accepted Manuscript

968

969 Avigad, D., Ziv, A. & Garfunkel, Z. (2001). Ductile and brittle shortening, extension-parallel

970 folds and maintenance of crustal thickness in the central Aegean (Cyclades, Greece).

971 Tectonics 20, 277–287.

Page 40 of 86 972

973 Baldwin, S. L. & Lister, G. S. (1998). Thermochronology of the South Cyclades Shear Zone,

974 Ios, Greece: effects of ductile shear in the argon partial retention zone. Journal of Geophysical

975 Research 103, 7315–7336.

976

977 Beyssac, O., Goffé, B., Chopin, C. & Rouzaud, J. N. (2002). Raman spectra of carbonaceous

978 material in metasediments: a new geothermometer. Journal of Metamorphic Geology 20, 859–

979 871.

980

981 Blake, M. C., Bonneau, M., Geyssant, J., Kienast, J. R., Lepvrier, C., Maluski, H. &

982 Papanikolaou, D. (1981). A geologic reconnaissance of the Cycladic blueschist belt, Greece.

983 Geological Society of America Bulletin 92, 247–254.

984

985 Bolhar, R., Ring, U. & Allen, C. M. (2010). An integrated zircon geochronological and

986 geochemical investigation into the Miocene plutonic evolution of the Cyclades, Aegean Sea,

987 Greece: Part 1: Geochronology. Contributions to Mineralogy and Petrology 160, 719–742.

988

989 Bonneau, M. (1984). Correlation of the Hellenic nappes in the south-east Aegean and their

990 tectonic reconstruction. In: Dixon, J. E., Robertson, A. H. F. (Eds.),The Geological Evolution

991 of the Eastern Mediterranean. Geological Society, London, Special Publications 17, 517–527.

992 Accepted Manuscript

993 Bonneau, M. & Kienast, J. R. (1982). Subduction, collision et schistes bleus: exemple de

994 l’Egée, Grèce. Bulletin de la Societe Geologique de France 24, 785–791.

995

Page 41 of 86 996 Bozkurt, E. (2007). Extensional v. contractional origin for the southern Menderes shear zone,

997 SW Turkey: tectonic and metamorphic implications. Geological Magazine 144, 191–210.

998

999 Bozkurt, E. & Park, R. G. (1997). Evolution of a mid-Tertiary extensional shear zone in the

1000 southern Menderes massif, western Turkey. Bulletin de la Societe Geologique de France 168,

1001 3–14.

1002

1003 Bozkurt, E., Satir, M. & Bugdaycioglu, C. (2011). Surprisingly young Rb/Sr ages from the

1004 Simav extensional detachment fault zone, northern Menderes Massif, Turkey. Journal of

1005 Geodynamics 52, 406–431.

1006

1007 Brandon, M. T., Roden-Tice, M. K. & Garver, J. I. (1998). Late Cenozoic exhumation of the

1008 Cascadia accretionary wedge in the Olympic Mountains, northwest Washington State.

1009 Geological Society of America Bulletin 110, 985–1009.

1010

1011 Brichau, S., Ring, U., Carter, A., Bolhar, R., Monié, P., Stockli, D. & Brunel, M. (2008).

1012 Timing, slip rate, displacement and cooling history of the Mykonos detachment footwall,

1013 Cyclades, Greece, and implications for the opening of the Aegean Sea basin. Journal of the

1014 Geological Society 165, 263–277.

1015

1016 Brichau, S., Ring,Accepted U., Carter, A., Monié, P., Bolhar, Manuscript R., Stockli, D. & Brunel, M. (2007).

1017 Extensional faulting on Tinos Island, Aegean Sea, Greece: How many detachments?

1018 Tectonics 26, TC4009.

1019

Page 42 of 86 1020 Brichau, S., Thomson, S. & Ring, U. (2010). Thermochronometric constraints on the tectonic

1021 evolution of the Serifos detachment, Aegean Sea, Greece. International Journal of Earth

1022 Sciences 99, 379–393.

1023

1024 Brunn, J. H., Argyriadis, I., Ricou, L. E., Poisson, A., Marcoux, J. & de Graciansky, P. C.

1025 (1976). Eléments majeurs de liaison entre Taurides et Hellénides. Bulletin de la Societe

1026 Geologique de France 18, 481–497.

1027

1028 Buick, I. S. (1991). The late Alpine evolution of an extensional shear zone, Naxos, Greece.

1029 Journal of the Geological Society, London 148, 93–103.

1030

1031 Buick, I. S. & Holland, T. J. B. (1989). The P-T-t path associated with crustal extension,

1032 Naxos, Cyclades, Greece. Geological Society, London, Special Publications 43, 365–369.

1033

1034 Cherniak, D. J. & Watson, E. B. (2001). Pb diffusion in zircon. Chemical Geology 172, 5–24.

1035

1036 Corti, G., Bonini, M., Conticelli, S., Innocenti, F., Manetti, P. & Sokoutis, D. (2003).

1037 Analogue modelling of continental extension: a review focused on the relations between the

1038 patterns of deformation and the presence of magma. Earth-Science Reviews 63, 169–247.

1039

1040 Crittenden, M. Accepted D., Coney, P. J. & Davis, G. H.Manuscript (1980). Cordilleran metamorphic core

1041 complexes. Geological Society of America Memoirs 153, 271–279.

1042

Page 43 of 86 1043 Cruz, M. J., Cunha, J. C., Merlet, C. & Sabaté, P. (1996). Dataçao pontual das monazitas da

1044 regiao de Itambe, Bahia, através da microssonda electrônica. XXXIX Congresso Brasileiro de

1045 Geologia, pp. 206–209.

1046

1047 Davis, G. H. & Coney, P. J. (1979). Geologic development of the Cordilleran metamorphic

1048 core complexes. Geology 7, 120–124.

1049

1050 Delvaux, D. & Sperner, B. (2003). New aspects of tectonic stress inversion with reference to

1051 the TENSOR program. In: Nieuwland, D. A. (Ed.), New Insights into Structural Interpretation

1052 and Modelling. Geological Society, London, Special Publications 212, 75–100.

1053

1054 Denèle, Y., Lecomte, E., Jolivet, L., Lacombe, O., Labrousse, L., Huet, B. & Le Pourhiet, L.

1055 (2011). Granite intrusion in a metamorphic core complex: The example of the Mykonos

1056 laccolith (Cyclades, Greece). Tectonophysics 501, 52–70.

1057

1058 Dewey, J. F. (1988). Extensional collapse of orogens. Tectonics 7, 1123–1139.

1059

1060 Didier, A., Bosse, V., Boulvais, P., Bouloton, J., Paquette, J. L., Montel, J. M. & Devidal, J.

1061 L. (2013). Disturbance versus preservation of U-Th-Pb ages in monazite during fluid-rock

1062 interaction: textural, chemical and isotopic in situ study in microgranites (Velay Dome,

1063 France). ContributionsAccepted to Mineralogy and Petrology Manuscript 165, 1051–1072.

1064

1065 Didier, A., Bosse, V., Cherneva, Z., Gautier, P., Georgieva, M., Paquette, J. L. & Gerdjikov, I.

1066 (2014). Syn-deformation fluid-assisted growth of monazite during renewed high-grade

Page 44 of 86 1067 metamorphism in metapelites of the Central (, Greece). Chemical Geology

1068 381, 206–222.

1069

1070 Duchêne, S., Aïssa, R. & Vanderhaeghe, O. (2006). Pressure-Temperature-Time Evolution of

1071 Metamorphic Rocks from Naxos (Cyclades, Greece): Constraints from Thermobarometry and

1072 Rb/Sr dating. Geodinamica Acta 19, 299–313.

1073

1074 Faccenna, C., Piromallo, C., Crespo-Blanc, A., Jolivet, L. & Rossetti, F. (2004). Lateral slab

1075 deformation and the origin of the western Mediterranean arcs. Tectonics 23, TC1012.

1076

1077 Farley, K. A. (2000). Helium diffusion from apatite: General behavior as illustrated by

1078 Durango fluorapatite. Journal of Geophysical Research 105, 2903–2914.

1079

1080 Faure, M., Bonneau, M. & Pons, J. (1991). Ductile deformation and syntectonic granite

1081 emplacement during the late Miocene extension of the Aegea (greece). Bulletin de la Societe

1082 Geologique de France 162, 3–11.

1083

1084 Fletcher, I. R., McNaughton, N. J., Davis, W. J. & Rsmussen, B. (2010). Matrix effects and

1085 calibration limitations in ion probe U-Pb and Th-Pb dating of monazite. Chemical Geology

1086 270, 31–44.

1087 Accepted Manuscript

1088 Forster, M. & Lister, G. (2009). Core-complex-related extension of the Aegean lithosphere

1089 initiated at the Eocene-Oligocene transition. Journal of Geophysical Research 114, B02401.

1090

Page 45 of 86 1091 Fossen, H., Teyssier, C. & Whitney, D. L. (2013). Transtensional folding. Journal of

1092 Structural Geology 56, 89–102.

1093

1094 Gabalda, S., Beyssac, O., Jolivet, L., Agard, P. & Chopin, C. (2009). Thermal structure of a

1095 fossil subduction wedge in the Western Alps. Terra Nova 21, 28–34.

1096

1097 Gasquet, D. et al. (2010). Miocene to Messinian deformation and hydrothermalism in the

1098 Lauzière Massif (French Western Alps): New U-Th-Pb and Argon ages. Bulletin de la Societe

1099 Geologique de France 181, 227–241.

1100

1101 Gautier, P. & Brun, J.-P. (1994). Ductile crust exhumation and extensional detachments in the

1102 central Aegean (Cyclades and Evvia Islands). Geodinamica Acta 7, 57–85.

1103

1104 Gautier, P., Brun, J.-P. & Jolivet, L. (1993). Structure and kinematics of upper cenozoic

1105 extensional detachment on Naxos and Paros (cyclades islands, Greece). Tectonics 12, 1180–

1106 1194.

1107

1108 Gessner, K., Collins, A. S., Ring, U. & Güngör, T. (2004). Structural and thermal history of

1109 poly-orogenic basement: U-Pb geochronology of granitoid rocks in the southern Menderes

1110 Massif, Western Turkey. Journal of the Geological Society, London 161, 93–101.

1111 Accepted Manuscript

1112 Gessner, K., Gallardo, L. A., Markwitz, V., Ring, U. & Thomson, S. N. (2013). What caused

1113 the denudation of the Menderes Massif: Review of crustal evolution, lithosphere structure,

1114 and dynamic topography in southwest Turkey. Gondwana Research 24, 243–274.

1115

Page 46 of 86 1116 Gessner, K., Ring, U., Johnson, C., Hetzel, R., Passchier, C. W. & Güngör, T. (2001). An

1117 active bivergent rolling-hinge detachment system: Central Menderes metamorphic core

1118 complex in western Turkey. Geology 29, 611–614.

1119

1120 Glodny, J. & Hetzel, R. (2007). Precise U–Pb ages of syn-extensional Miocene intrusions in

1121 the central Menderes Massif, western Turkey. Geological Magazine 144, 235.

1122

1123 Grasemann, B. & Petrakakis, K. (2007). Evolution of the Serifos Metamorphic Core

1124 Complex. In: Lister, G., Foster, M., Ring, U. (Eds.), Inside the Aegean Core Complexes.

1125 Journal of the Virtual Explorer 27.

1126

1127 Grasemann, B., Schneider, D. A., Stöckli, D. F. & Iglseder, C. (2012). Miocene bivergent

1128 crustal extension in the Aegean: Evidence from the western Cyclades (Greece). Lithosphere 4,

1129 23–39.

1130

1131 Groppo, C., Forster, M., Lister, G. & Compagnoni, R. (2009). Glaucophane schists and

1132 associated rocks from Sifnos (Cyclades, Greece): New constraints on the P–T evolution from

1133 oxidized systems. Lithos 109, 254–273.

1134

1135 Grove, M. & Harrison, T. M. (1996). Ar-40* diffusion in Fe-rich biotite. American

1136 Mineralogist 81,Accepted 940–951. Manuscript

1137

1138 Harrison, T. M. (1981). Diffusion of 40Ar in hornblende. Contributions to Mineralogy and

1139 Petrology 78, 324–331.

1140

Page 47 of 86 1141 Harrison, T. M., Célérier, J., Aikman, A. B., Hermann, J. & Heizler, M. T. (2009). Diffusion

1142 of 40Ar in muscovite. Geochimica et Cosmochimica Acta 73, 1039–1051.

1143

1144 Henjes-Kunst, F. & Kreuzer, H. (1982). Isotopic dating of pre-Alpidic rocks from the island

1145 of Ios (Cyclades, Greece). Contributions to Mineralogy and Petrology 80, 245–253.

1146

1147 Hetzel, R., Passchier, C. W., Ring, U. & Dora, Ö. O. (1995a). Bivergent extension in orogenic

1148 belts: the Menderes massif (southwestern Turkey). Geology 23, 455–458.

1149

1150 Hetzel, R., Ring, U., Akal, C. & Troesch, M. (1995b). Miocene NNE-directed extensional

1151 unroofing in the Menderes Massif, southwestern Turkey. Journal of the Geological Society,

1152 London 152, 639–654.

1153

1154 Hezel, D. C., Kalt, A., Marschall, H. R., Ludwig, T. & Meyer, H.-P. (2011). Major-element

1155 and Li, Be compositional evolution of tourmaline in an S-type granite-pegmatite system and

1156 its country rocks: an example from Ikaria, Aegean Sea, Greece. The Canadian Mineralogist

1157 49, 321–340.

1158

1159 Huet, B., Labrousse, L. & Jolivet, L. (2009). Thrust or detachment? Exhumation processes in

1160 the Aegean: Insight from a field study on Ios (Cyclades, Greece). Tectonics 28, TC3007.

1161 Accepted Manuscript

1162 Iglseder, C., Grasemann, B., Rice, A. H. N., Petrakakis, K. & Schneider, D. A. (2011).

1163 Miocene south directed low-angle normal fault evolution on Kea Island (West Cycladic

1164 Detachment System, Greece). Tectonics 30, TC4013.

1165

Page 48 of 86 1166 Iglseder, C., Grasemann, B., Schneider, D. A., Petrakakis, K., Miller, C., Klötzli, U. S., Thöni,

1167 M., Zámolyi, A. & Rambousek, C. (2009). I and S-type plutonism on Serifos (W-Cyclades,

1168 Greece). Tectonophysics 473, 69–83.

1169

1170 Imber, J., Holdsworth, R. E., Butler, C. A. & Strachan, R. A. (2001). A reappraisal of the

1171 Sibson-Scholz fault zone model: The nature of the frictional to viscous (“brittle-ductile”)

1172 transition along a long-lived, crustal-scale fault, Outer Hebrides, Scotland. Tectonics 20, 601–

1173 624.

1174

1175 Jackson, S. E., Pearson, N. J., Griffin, W. L. & Belousova, E. A. (2004). The application of

1176 laser ablation-inductively coupled plasma-mass spectrometry to in situ U-Pb zircon

1177 geochronology. Chemical Geology 211, 47–69.

1178

1179 Janots, E., Berger, A. & Gnos, E. (2012). Constraints on fluid evolution during

1180 metamorphism from U-Th-Pb systematics in Alpine hydrothermal monazite. Chemical

1181 Geology 326-327, 61–71.

1182

1183 Jansen, J. B. H. & Schuiling, R. D. (1976). Metamorphism on Naxos-Petrology and

1184 geothermal gradients. American Journal of Science 276, 1225–1253.

1185

1186 Jolivet, L. et al. Accepted(2013). Aegean tectonics: Strain localisation,Manuscript slab tearing and trench retreat.

1187 Tectonophysics 597, 1–33.

1188

Page 49 of 86 1189 Jolivet, L., Augier, R., Faccenna, C., Negro, F., Rimmelé, G., Agard, P., Robin, C., Rossetti,

1190 F. & Crespo-Blanc, A. (2008). Subduction, convergence and the mode of backarc extension in

1191 the Mediterranean region. Bulletin de la Societe Geologique de France 179, 525–550.

1192

1193 Jolivet, L. & Brun, J.-P. (2010). Cenozoic geodynamic evolution of the Aegean. International

1194 Journal of Earth Sciences 99, 109–138.

1195

1196 Jolivet, L. & Faccenna, C. (2000). Mediterranean extension and the Africa-Eurasia collision.

1197 Tectonics 19, 1095–1106.

1198

1199 Jolivet, L., Faccenna, C. & Piromallo, C. (2009). From mantle to crust: Stretching the

1200 Mediterranean. Earth and Planetary Science Letters 285, 198–209.

1201

1202 Jolivet, L., Famin, V., Mehl, C., Parra, T., Aubourg, C., Hébert, R. & Philippot, P. (2004a).

1203 Progressive strain localisation, boudinage and extensional metamorphic complexes, the

1204 Aegean Sea case. In: Whitney, D. L., Teyssier, C., Siddoway, C. S. (Eds.), Gneiss Domes in

1205 Orogeny. Geological Society of America Special Papers 380, 185–210.

1206

1207 Jolivet, L., Lecomte, E., Huet, B., Denèle, Y., Lacombe, O., Labrousse, L., Le Pourhiet, L. &

1208 Mehl, C. (2010). The North Cycladic Detachment System. Earth and Planetary Science

1209 Letters 289, 87–Accepted104. Manuscript

1210

1211 Jolivet, L. & Patriat, M. (1999). Ductile extension and the formation of the Aegean Sea. In:

1212 Durand, B., Jolivet, L., Horvàth, F., Séranne, M. (Eds.), The Mediterranean Basins: Tertiary

Page 50 of 86 1213 Extension Within the Alpine Orogen. Geological Society, London, Special Publications 156,

1214 427–456.

1215

1216 Jolivet, L., Rimmelé, G., Oberhänsli, R., Goffé, B. & Candan, O. (2004b). Correlation of syn-

1217 orogenic tectonic and metamorphic events in the Cyclades, the Lycian nappes and the

1218 Menderes massif. Geodynamic implications. Bulletin de la Societe Geologique de France 175,

1219 217–238.

1220

1221 Katzir, Y., Matthews, A., Garfunkel, Z., Schliestedt, M. & Avigad, D. (1996). The tectono-

1222 metamorphic evolution of a dismembered ophiolite (Tinos, Cyclades, Greece). Geological

1223 Magazine 133, 237–254.

1224

1225 Keay, S. & Lister, G. (2002). African provenance for the metasediments and metaigneous

1226 rocks of the Cyclades, Aegean Sea, Greece. Geology 30, 235–238.

1227

1228 Keay, S., Lister, G. & Buick, I. (2001). The timing of partial melting, Barrovian

1229 metamorphism and granite intrusion in the Naxos metamorphic core complex, Cyclades,

1230 Aegean Sea, Greece. Tectonophysics 342, 275–312.

1231

1232 Keiter, M., Piepjohn, K., Ballhaus, C., Lagos, M. & Bode, M. (2004). Structural development

1233 of high-pressureAccepted metamorphic rocks on Syros island Manuscript (Cyclades, Greece). Journal of Structural

1234 Geology 26, 1433–1445.

1235

Page 51 of 86 1236 Ketcham, R. A., Donelick, R. A. & Carlson, W. D. (1999). Variability of apatite fission-track

1237 annealing kinetics: III. Extrapolation to geological time scales. American Mineralogist 84,

1238 1235–1255.

1239

1240 Kokkalas, S. & Aydin, A. (2013). Is there a link between faulting and magmatism in the

1241 south-central Aegean Sea? Geological Magazine 150, 193–224.

1242

1243 Kruckenberg, S. C., Vanderhaeghe, O., Ferré, E. C., Teyssier, C. & Whitney, D. L. (2011).

1244 Flow of partially molten crust and the internal dynamics of a migmatite dome, Naxos, Greece.

1245 Tectonics 30, TC3001.

1246

1247 Kuhlemann, J., Frisch, W., Dunkl, I., Kázmér, M. & Schmiedl, G. (2004). Miocene

1248 siliciclastic deposits of Naxos Island: geodynamic and environmental implications for the

1249 evolution of the southern Aegean Sea (Greece). In: Bernet, M., Spiegel, C. (Eds.), Detrital

1250 Thermochronology-Provenance Analysis, Exhumation, and Landscape Evolution of Mountain

1251 Belts. Geological Society of America Special Papers 378, 51–65.

1252

1253 Kumerics, C., Ring, U., Brichau, S., Glodny, J. & Monié, P. (2005). The extensional Messaria

1254 shear zone and associated brittle detachment faults, Aegean Sea, Greece. Journal of the

1255 Geological Society 162, 701–721.

1256 Accepted Manuscript

1257 Lahfid, A., Beyssac, O., Deville, E., Negro, F., Chopin, C. & Goffé, B. (2010). Evolution of

1258 the Raman spectrum of carbonaceous material in low-grade metasediments of the Glarus Alps

1259 (Switzerland). Terra Nova 22, 354–360.

1260

Page 52 of 86 1261 Laurent, V., Beaudoin, A., Jolivet, L., Arbaret, L., Augier, R., Rabillard, A. & Menant, A.

1262 (2015). Interrelations between extensional shear zones and synkinematic intrusions: The

1263 example of Ikaria Island (NE Cyclades, Greece). Tectonophysics 651–652, 152–171.

1264

1265 Lecomte, E., Jolivet, L., Lacombe, O., Denèle, Y., Labrousse, L. & Le Pourhiet, L. (2010).

1266 Geometry and kinematics of Mykonos detachment, Cyclades, Greece: Evidence for slip at

1267 shallow dip. Tectonics 29, TC5012.

1268

1269 Lee, J. & Lister, G. S. (1992). Late Miocene ductile extension and detachment faulting,

1270 Mykonos, Greece. Geology 20, 121–124.

1271

1272 Le Pichon, X. & Angelier, J. (1979). The Hellenic arc and trench system: a key to the

1273 neotectonic evolution of the eastern Mediterranean area. Tectonophysics 60, 1–42.

1274

1275 Le Pichon, X. & Kreemer, C. (2010). The Miocene-to-present kinematic evolution of the

1276 eastern mediterranean and Middle East and its implications for dynamics. Annual Review of

1277 Earth and Planetary Sciences 38, 323–351.

1278

1279 Le Pourhiet, L., Huet, B., May, D. A., Labrousse, L. & Jolivet, L. (2012). Kinematic

1280 interpretation of the 3D shapes of metamorphic core complexes. Geochemistry Geophysics

1281 Geosystems 13, AcceptedQ09002. Manuscript

1282

1283 Lips, A. L. W., Cassard, D., Sözbilir, H., Yilmaz, H. & Wijbrans, J. R. (2001). Multistage

1284 exhumation of the Menderes Massif, western Anatolia (Turkey). International Journal of

1285 Earth Sciences 89, 781–792.

Page 53 of 86 1286

1287 Lister, G. S., Banga, G. & Feenstra, A. (1984). Metamorphic core complexes of Cordilleran

1288 type in the Cyclades, Aegean Sea, Greece. Geology 12, 221–225.

1289

1290 Ludwig, K. R. (2001). Isoplot/Ex rev. 2.49 - A Geochronological Toolkit for Microsoft Excel.

1291 Berkeley Geochronology canter. Special publication, No. 1a.

1292

1293 Malinverno, A. & Ryan, W. B. F. (1986). Extension in the Tyrrhenian Sea and shortening in

1294 the Apennines as result of arc migration driven by sinking of the lithosphere. Tectonics 5,

1295 227–245.

1296

1297 Martin, L. A. J., Ballèvre, M., Boulvais, P., Halfpenny, A., Vanderhaeghe, O., Duchêne, S. &

1298 Deloule, E. (2011). Garnet re-equilibration by coupled dissolution-reprecipitation: evidence

1299 from textural, major element and oxygen isotope zoning of “cloudy” garnet: Journal of

1300 Metamorphic Geology 29, 213–231.

1301

1302 Mehl, C., Jolivet, L. & Lacombe, O. (2005). From ductile to brittle: Evolution and

1303 localization of deformation below a crustal detachment (Tinos, Cyclades, Greece). Tectonics

1304 24, TC4017.

1305

1306 Mehl, C., Jolivet,Accepted L., Lacombe, O., Labrousse, L. &Manuscript Rimmele, G. (2007). Structural evolution

1307 of Andros island (Cyclades, Greece): a key to the behaviour of a flat detachment within an

1308 extending continental crust. In: Taymaz, T., Dilek, Y., Ylmaz, Y. (Eds.), The Geodynamics of

1309 the Aegean and Anatolia. Geological Society, London, Special Publications 291, 41–73.

1310

Page 54 of 86 1311 Melinte-Dobrinescu, M. C. et al. (2009). The Messinian Salinity Crisis in the Dardanelles

1312 region: Chronostratigraphic constraints. Palaeogeography, Palaeoclimatology, Palaeoecology

1313 278, 24–39.

1314

1315 Menant, A., Jolivet, L., Augier, R. & Skarpelis, N. (2013). The North Cycladic Detachment

1316 System and associated mineralization, Mykonos, Greece: Insights on the evolution of the

1317 Aegean domain. Tectonics 32, 433–452.

1318

1319 Norton, M. G. (1986). Late Caledonide Extension in western Norway: A response to extreme

1320 crustal thickening. Tectonics 5, 195–204.

1321

1322 Oberhansli, R., Monie, P., Candan, O., Warkus, F. C., Partzsch, J. H. & Dora, O. O. (1998).

1323 The age of blueschist metamorphism in the Mesozoic cover series of the Menderes Massif.

1324 Schweizerische Mineralogische Und Petrographische Mitteilungen 78, 309–316.

1325

1326 Papanikolaou, D. (1978). Contribution to the geology of Ikaria island, Aegean sea. Annales

1327 Géologiques des Pays Helléniques 29, 1–28.

1328

1329 Paquette, J. L. & Tiepolo, M. (2007). High resolution (5 μm) U-Th-Pb isotope dating of

1330 monazite with excimer laser ablation (ELA)-ICP-MS. Chemical Geology 240, 222–237.

1331 Accepted Manuscript

1332 Parra, T., Vidal, O. & Jolivet, L. (2002). Relation between the intensity of deformation and

1333 retrogression in blueschist metapelites of Tinos Island (Greece) evidenced by chlorite-mica

1334 local equilibria. Lithos 63, 41–66.

1335

Page 55 of 86 1336 Pasteris, J. D. & Wopenka, B. (1991). Raman-Spectra of Graphite as Indicators of Degree of

1337 Metamorphism. Canadian Mineralogist 29, 1–9.

1338

1339 Pe-Piper, G. & Photiades, A. (2006). Geochemical characteristics of the Cretaceous ophiolitic

1340 rocks of Ikaria island, Greece. Geological Magazine 143, 417–429.

1341

1342 Photiades, A. D. (2002a). The ophiolitic molasse unit of Ikaria Island (Greece). Turkish

1343 Journal of Earth Sciences 11, 27–38.

1344

1345 Photiades, A. D. (2002b). Geological Map of Greece Scale, 1:50000: Island of Ikaria. Institute

1346 of Geology and Mineral Exploration.

1347

1348 Platt, J. P., Behr, W. M. & Cooper, F. J. (2014). Metamorphic core complexes: windows into

1349 the mechanics and rheology of the crust. Journal of the Geological Society.

1350

1351 Platt, J. P. & Vissers, R. L. M. (1989). Extensional collapse of thickened continental

1352 lithosphere: A working hypothesis for the Alboran Sea and Gibraltar arc. Geology 17, 540–

1353 543.

1354

1355 Pourteau, A., Candan, O. & Oberhänsli, R. (2010). High-pressure metasediments in central

1356 Turkey: ConstraintsAccepted on the Neotethyan closure history. Manuscript Tectonics 29, TC5004.

1357

1358 Pourteau, A., Sudo, M., Candan, O., Lanari, P., Vidal, O. & Oberhänsli, R. (2013). Neotethys

1359 closure history of Anatolia: insights from 40Ar-39Ar geochronology and P-T estimation in

1360 high-pressure metasedimentary rocks. Journal of Metamorphic Geology 31, 585–606.

Page 56 of 86 1361

1362 Rimmelé, G., Jolivet, L., Oberhänsli, R. & Goffé, B. (2003b). Deformation history of the

1363 high-pressure Lycian Nappes and implications for tectonic evolution of SW Turkey.

1364 Tectonics 22, 1007.

1365

1366 Rimmelé, G., Oberhänsli, R., Goffé, B., Jolivet, L., Candan, O. & Cetinkaplan, M. (2003a).

1367 First evidence of high-pressure metamorphism in the “Cover Series” of the southern

1368 Menderes Massif. Tectonic and metamorphic implications for the evolution of SW Turkey.

1369 Lithos 71, 19–46.

1370

1371 Ring, U. (2007). The Geology of Ikaria Island: The Messaria extensional shear zone, granites

1372 and the exotic Ikaria nappe. In: Lister, G., Foster, M., Ring, U. (Eds.), Inside the Aegean Core

1373 Complexes. Journal of the Virtual Explorer 27.

1374

1375 Ring, U. et al. (2007). Early exhumation of high-pressure rocks in extrusion wedges: Cycladic

1376 blueschist unit in the eastern Aegean, Greece, and Turkey. Tectonics 26, TC2001.

1377

1378 Ring, U. & Collins, A. S. (2005). U-Pb SIMS dating of synkinematic granites: timing of core-

1379 complex formation in the northern Anatolide belt of western Turkey. Journal of the

1380 Geological Society 162, 289–298.

1381 Accepted Manuscript

1382 Ring, U., Glodny, J., Will, T. M. & Thomson, S. (2011). Normal faulting on Sifnos and the

1383 South Cycladic Detachment System, Aegean Sea, Greece. Journal of the Geological Society

1384 168, 751–768.

1385

Page 57 of 86 1386 Ring, U., Glodny, J., Will, T. & Thomson, S. (2010). The Hellenic Subduction System: High-

1387 Pressure Metamorphism, Exhumation, Normal Faulting, and Large-Scale Extension. In:

1388 Jeanloz, R. & Freeman, K. H. (eds) Annual Review of Earth and Planetary Sciences. Palo

1389 Alto: Annual Reviews, 45–76.

1390

1391 Ring, U., Laws, S. & Bernet, M. (1999). Structural analysis of a complex nappe sequence and

1392 late-orogenic basins from the Aegean Island of Samos, Greece. Journal of Structural Geology

1393 21, 1575–1601.

1394

1395 Rosenbaum, G., Lister, G. S. & Duboz, C. (2002). Relative motions of Africa, Iberia and

1396 Europe during Alpine orogeny. Tectonophysics 359, 117–129.

1397

1398 Rosenbaum, G., Ring, U. & Kühn, A. (2007). Tectonometamorphic evolution of high-

1399 pressure rocks from the island of Amorgos (Central Aegean, Greece). Journal of the

1400 Geological Society, London 164, 425–438.

1401

1402 Royden, L. H. (1993). Evolution of retreating subduction boundaries formed during

1403 continental collision. Tectonics 12, 629–638.

1404

1405 Sanchez-Gomez, M., Avigad, D. & Heimann, A. (2002). Geochronology of clasts in

1406 allochthonous MioceneAccepted sedimentary sequences on ManuscriptMykonos and Paros Islands: implications

1407 for back-arc extension in the Aegean Sea. Journal of the Geological Society, London 159, 45–

1408 60.

1409

Page 58 of 86 1410 Schärer, U. (1984). The effect of initial 230Th disequilibrium on young U-Pb ages: the

1411 Makalu case, Himalaya. Earth and Planetary Science Letters 67, 191–204.

1412

1413 Sengör, A. M. C., Tüysüz, O., İmren, C., Sakınç, M., Eyidoğan, H., Görür, N., Le Pichon, X.

1414 & Rangin, C. (2005). The North Anatolian Fault: A New Look. Annual Review of Earth and

1415 Planetary Sciences 33, 37–112.

1416

1417 Seydoux-Guillaume, A. M., Paquette, J. L., Wiedenbeck, M., Montel, J. M. & Heinrich, W.

1418 (2002). Experimental resetting of the U-Th-Pb systems in monazite. Chemical Geology 240,

1419 222–237.

1420

1421 Steck, A. & Hunziker, J. (1994). The Tertiary structural and thermal evolution of the Central

1422 Alps-compressional and extensional structures in an orogenic belt. Tectonophysics 238, 229–

1423 254.

1424

1425 Sternai, P., Jolivet, L., Menant, A. & Gerya, T. (2014). Driving the upper plate surface

1426 deformation by slab rollback and mantle flow. Earth and Planetary Science Letters 405, 110–

1427 118.

1428

1429 Stöckhert, B., Brix, M. R., Kleinschrodt, R., Hurford, A. J. & Wirth, R. (1999).

1430 ThermochronometryAccepted and microstructures of quartz -aManuscript comparison with experimental flow laws

1431 and predictions on the temperature of the brittle-plastic transition. Journal of Structural

1432 Geology 21, 351–369.

1433

Page 59 of 86 1434 Thomson, S. N., Ring, U., Brichau, S., Glodny, J. & Will, T. M. (2009). Timing and nature of

1435 formation of the Ios metamorphic core complex, southern Cyclades, Greece. In: Ring, U.,

1436 Wernicke, B. (Eds.), Extending a Continent: Architecture, Rheology and Heat Budget.

1437 Geological Society, London, Special Publications 321, 139–167.

1438

1439 Tirel, C., Brun, J.-P. & Burov, E. (2008). Dynamics and structural development of

1440 metamorphic core complexes. Journal of Geophysical Research 113, B04403.

1441

1442 Tomascak, P. B., Krogstad, E. J. & Walker, R. J. (1996). U-Pb monazite geochronology of

1443 granitic rocks from Maine: Implications for Late Paleozoic tectonics in the northern

1444 Appalachians. The Journal of Geology 104, 185–195.

1445

1446 Trotet, F., Jolivet, L. & Vidal, O. (2001a). Tectono-metamorphic evolution of Syros and

1447 Sifnos islands (Cyclades, Greece). Tectonophysics 338, 179–206.

1448

1449 Trotet, F., Vidal, O. & Jolivet, L. (2001b). Exhumation of Syros and Sifnos metamorphic

1450 rocks (Cyclades, Greece). New constraints on the P-T paths. European Journal of Mineralogy

1451 13, 901–920.

1452

1453 Tschegg, C. & Grasemann, B. (2009). Deformation and alteration of a granodiorite during

1454 low-angle normalAccepted faulting (Serifos, Greece). Lithosphere Manuscript 1, 139–154.

1455

1456 Urai, J. L., Schuiling, R. D. & Jansen, J. B. H. (1990). Alpine deformation on Naxos (Greece),

1457 in: Knipe, R. J., Rutter, E. H. (Eds.), Deformation Mechanisms, Rheology and Tectonics.

1458 Geological Society, London, Special Publications 54, 509–522.

Page 60 of 86 1459

1460 van Achterbergh, E., Ryan, C. G., Jackson, S. E. & Griffin, W. L. (2001). Data reduction

1461 software for LA-ICP-MS. In: Sylvester, P. (Ed.), Laser ablation-ICP-MS in the Earth Science.

1462 Mineralogical Association of Canada 29, 239–243.

1463

1464 Vandenberg, L. C. & Lister, G. S. (1996). Structural analysis of basement tectonites from the

1465 Aegean metamorphic core complex of Ios, Cyclades, Greece. Journal of Structural Geology

1466 18, 1437–1454.

1467

1468 Vanderhaeghe, O. (2004). Structural development of the Naxos migmatite dome. In: Whitney,

1469 D. L., Teyssier, C., Siddoway, C. S. (eds.), Gneiss Domes in Orogeny. Geological Society of

1470 America Special Papers 211–227.

1471

1472 van der Maar, P. A., Feenstra, A., Manders, B. & Jansen, J. B. H. (1981). The petrology of the

1473 island of Sikinos, Cyclades, Greece, in comparison with that of the adjacent island of Ios.

1474 Neues Jahrbuch für Mineralogie-Abhandlungen 10, 459–469.

1475

1476 van Hinsbergen, D. J. J. (2010). A key extensional metamorphic complex reviewed and

1477 restored: The Menderes Massif of western Turkey. Earth-Science Reviews 102, 60–76.

1478

1479 van Hinsbergen,Accepted D. J. J., Hafkenscheid, E., Spakman, Manuscript W., Meulenkamp, J. E. & Wortel, R.

1480 (2005). Nappe stacking resulting from subduction of oceanic and continetal lithosphere below

1481 Greece. Geology 33, 325–328.

1482

1483 Villa, I. M. (1998). Isotopic closure. Terra Nova 10, 42–47.

Page 61 of 86 1484

1485 Vitale Brovarone, A., Beyssac, O., Malavieille, J., Molli, G., Beltrando, M. & Compagnoni,

1486 R. (2013). Stacking and metamorphism of continuous segments of subducted lithosphere in a

1487 high-pressure wedge: The example of Alpine Corsica (France). Earth-Science Reviews 116,

1488 35–56.

1489

1490 Weinberg, R. F. & Hasalová, P. (2015). Water-fluxed melting of the continental crust: A

1491 review. Lithos 212–215, 158–188.

1492

1493 Wernicke, B. (1981). Low-angle normal faults in the Basin and Range Province: nappe

1494 tectonics in an extending orogen. Nature 291, 645–648.

1495

1496 Wernicke, B. (1985). Structural discordance between Neogene detachments and frontal sevier

1497 thrusts, central Mormon Mountain, southern Nevada. Tectonics 4, 213–246.

1498

1499 Whitney, D. L. & Evans, B. W. (2010). Abbreviations for names of rock-forming minerals.

1500 American Mineralogist 95, 185–187.

1501

1502 Wijbrans, J. R., Schliestedt, M. & York, D. (1990). Single grain argon laser probe dating of

1503 phengites from the blueschist to greenschist transition on Sifnos (Cyclades, Greece).

1504 Contributions toAccepted Mineralogy and Petrology 104, 582 Manuscript–593.

1505

1506 Will, T., Okrusch, M., Schmädicke, E. & Chen, G. L. (1998). Phase relations in the

1507 greenschist-blueschist-amphibolite-eclogite facies in the system Na2O-CaO-FeO-MgO-

Page 62 of 86 1508 Al2O3-SiO2-H2O (NCFMASH), with application to metamorphic rocks from Samos, Greece.

1509 Contributions to Mineralogy and Petrology 132, 85–102.

1510

1511 Table caption

1512

1513 Table 1: RSCM peak temperature results

1514 RSCM results classified by increasing temperatures. For each sample are indicated the GPS

1515 position, the total number of Raman spectra (n) performed, the mean R2 ratio and temperature

1516 calculated, both associated with their standard deviation (SD) related to the intra-sample

1517 heterogeneity. Location of samples and Tmax results are given on Fig. 12.

1518

1519

1520 Figure captions

1521

1522 Figure 1: Tectonic map of the Aegean domain

1523 Tectonic map of the Aegean domain showing the main geological units and structures related

1524 to both synorogenic and postorogenic episodes, modified after Jolivet et al. (2013) and

1525 references therein. Original map has been modified incorporating recent works (Ring et al.,

1526 1999; Kumerics et al., 2005; Rosenbaum et al., 2007; Huet et al., 2009; Grasemann et al.,

1527 2012; Augier et al., 2015). NCDS: North Cycladic Detachment System; WCDS: West

1528 Cycladic DetachmentAccepted System; NDP: Naxos-Paros Detachment;Manuscript SCT: South Cycladic Thrust.

1529

1530 Figure 2: Geological map of Ikaria

1531 (a) New geological map of Ikaria proposed in this study. Lithologic outlines correspond to

1532 new field observations and a compilation of existing maps (e.g. Papanikolaou, 1978;

Page 63 of 86 1533 Photiades, 2002b; Kumerics et al., 2005; Ring, 2007). Also indicated are two representative

1534 tectonometamorphic piles for the eastern and the western parts of Ikaria.

1535

1536 Figure 3: New geological features and precisions about nature of contacts

1537 (a) The intrusive contact between Raches granite and marbles of Ikaria unit is clearly marked

1538 by the presence of granitic dykes within marbles. (b) Southward extension of the Fanari

1539 detachment at Therma, separating conglomerates of Fanari unit and metasediments of Agios-

1540 Kirykos unit. (c) Panorama of the Fanari detachment at Gialiskari between sediments of

1541 Fanari unit and the Raches granite. (d) Close-up view of granite-derived cataclasites beneath

1542 the Fanari detachment at Gialiskari. (e) Details of migmatites from the lower parts of Ikaria

1543 unit preserved in the south of Raches granite. See Fig. 2 for location.

1544

1545 Figure 4: Main planar fabrics

1546 (a) Foliation-map of Ikaria. Geometry of sedimentary bedding in Fanari unit is also showed.

1547 (b) Statistics of the main foliation geometry. Poles of foliation are presented in Schmidt’s

1548 lower hemisphere equal-area projection and preferred orientations of foliation is given by the

1549 rose-diagram. The elongation of the cloud allows retrieving the geometry of the dome axis.

1550 Also note that the asymmetry of the cloud calls for the asymmetry of the dome with a steeper

1551 southeast flank. (c) Simplified geological map showing foliation trajectories and traces of the

1552 main shear zones.

1553 Accepted Manuscript

1554 Figure 5: Stretching lineation

1555 (a) Stretching lineation-map of Ikaria. Note that the trend of the lineation shows very little

1556 dispersion. It is noteworthy that the strike of the stretching lineation describes slightly curved

1557 patterns from N-S to more NNE-SSW directions. Indicated are the results of the fault-slip data

Page 64 of 86 1558 inversion. (b) Statistics on the preferred orientation of stretching lineation presented in

1559 Schmidt’s lower hemisphere equal-area projection.

1560

1561 Figure 6: Upward strain gradient within Ikaria unit

1562 (a) Incomplete transposition of the compositional layering into a main, flat-lying penetrative

1563 foliation within the deepest parts of the structure. (b) More advanced transposition of the

1564 compositional layering into the flat-lying penetrative foliation. Note that boudin-shaped

1565 volume of rocks preserved isoclinal folds between decameter-scale top-to-the-N shear bands.

1566 (c) Close-up view of the cross-cutting relationships compositional layering and the main

1567 foliation. Note that fold axes are now mostly parallel to the stretching direction. (d) Mylonitic

1568 deformation of the upper parts of Ikaria unit. The foliation dips gently toward the north.

1569 Metabasites form asymmetric pinch-and-swell boudinage indicating top-to-the-N kinematics.

1570 (e) Typical mylonites from the uppermost parts of Ikaria unit, near Evdilos. Quartz veins have

1571 been transposed and stretched into the shear direction and forming asymmetric sigmoids

1572 indicating top-to-the-NNE kinematics. (f) Composite cross section showing the first-order

1573 strain gradient and the structural position of outcrops of Fig. 6. Trace is represented on Fig. 2.

1574

1575 Figure 7: Physical conditions of the deformation within Ikaria unit

1576 (a) Small-scale top-to-the-S shear bands operating in the lower parts of Ikaria unit. The

1577 volume of rocks involved in the deformation is quite large and the amphibolite-facies

1578 associations areAccepted quite well preserved. Note that Manuscript in these incipiently deformed rocks,

1579 deformation is rather symmetrical and local top-to-the-S kinematics are sometimes observed.

1580 (b) Close-up view of syn-greenschist-facies mylonites from the core of the Agios-Kirykos

1581 shear zone in the upper parts of Ikaria unit.

1582

Page 65 of 86 1583 Figure 8: Deformation in Agios-Kirykos unit

1584 Outcrop pictures of rather weak (a) and intense (b) top-to-the-NE asymmetric deformation in

1585 Agios-Kirykos unit. Note that shearing, developed in the greenschist-facies conditions,

1586 displays a clear evolution to more localized or even cataclastic flow. (c) Large-scale view of

1587 the core of the Fanari shear zone at Agios Kiriaki. Exposed are the mylonites of the

1588 uppermost parts of Ikaria unit, on the first plane, while sediments of Fanari unit crop out in

1589 the background. Close-up views of the deformation of (d) marble and (e) metapelite layers.

1590 Marble layers display successive stages of symmetric boudinage, ranging from ductile to

1591 strictly brittle while metapelite levels show a strong non-coaxial component consistent with

1592 an overall top-to-the-NE kinematics.

1593

1594 Figure 9: The Fanari detachment plane

1595 Large-scale representative view of the Fanari detachment plane at Fanari. The detachment

1596 plane is stripped of its sedimentary cover over several hundreds of square meters. Sediments

1597 that are almost vertical are preserved in the incised gullies and all along the coast from Fanari

1598 to . The plane carries series of large-scale corrugations and crescent-shaped

1599 structures indicating a consistent top-to-the-NNE kinematics. Note that the detachment plane

1600 is cut across by series of high-angle faults carrying sub-horizontal striations. Also are

1601 represented stereographic projections of striations and kinematics of both the detachment

1602 plane and the faults that offset the plane.

1603 Accepted Manuscript

1604 Figure 10: Using microstructures to unfold the Fanari detachment

1605 (a) Sketch depicting the probable geometry of the Fanari detachment prior to its tilting

1606 together with the whole system (see Fig. 9). The result is a system where the detachment

1607 plane operates at shallow angle with dextral-normal-sense. In the restored position, sediments

Page 66 of 86 1608 appear pervasively affected by a single set of conjugate normal faults. (b) Outcrop pictures

1609 and stereographic projections of fault systems before and after back-tilting rotation for sites 6

1610 and 7 (see Fig. 5 for location). In this restored position, paleostress solutions are all consistent

1611 with a unique and common NNE-SSW to NE-SW extensional stretching. This orientation is

1612 consistent with the paleostress field deduced from other stations throughout the island (see

1613 Fig. 11). Note that, conversely, three successive stress states are required to account for the

1614 heterogeneous fault populations present in their current position.

1615

1616 Figure 11: Small-scale brittle structures

1617 Detailed results of the fault-slip data inversion. Fault planes, associated striae and results of

1618 inversion were plotted using WinTensor software in Schmidt’s lower hemisphere equal-area

1619 projection (Delvaux and Sperner, 2003). Also is presented a representative outcrop recognized

1620 as demonstrative of a brittle stage subsequently developed after the ductile one (site 1; see

1621 location on Fig. 5). Faults occur as a dense array of rather steep normal faults. Note the

1622 consistency of the NNE-SSW to NE-SW direction of extension all over the island.

1623

1624 Figure 12: RSCM peak temperature results

1625 (a) Sampling map used for the RSCM study. In the bottom right corner, Tmax are sorted by

1626 increasing temperature, where errors brackets correspond to intra-sample heterogeneity

1627 standard deviation. Note that temperature is comparable for two samples, positioned side by

1628 side in the chart,Accepted for Ikaria unit on one hand, and forManuscript Agios Kirykos unit on the other hand.

1629 But the two units are characterized by clearly differentiable bulk temperature. (b) Tmax vs

1630 structural position for the northeastern part (samples used for (b) are written in bold on the

1631 map). Note the gap of temperature near the Agios-Kirykos shear zone, quite clear on the

1632 Therma cross section. (c) Increasing temperature approaching the Raches granite.

Page 67 of 86 1633

1634 Figure 13: U-Th-Pb analyses on monazite from migmatite sample IK01A

1635 (a) Typical textural relationships between monazite (Mnz) crystals and the magmatic

1636 paragenesis as explored by BSE mean. Mnz 3 is included into biotite (Bt). (b) Details of the

1637 internal texture of Mnz 3 and Mnz 4 monazite crystals. BSE image reveal a clear core-rim

1638 textures. Is also shown the location of ICPMS laser ablation analysis (9 µm) and their

1639 corresponding 208Pb/232Th ages (2σ level). Note the correlation between ages and the core-rim

1640 textures. (c) 206Pb/238U vs 208Pb/232Th diagram for all data showing a 208Pb/232Th age of 15.7 ±

1641 0.2 Ma concordant with the 206Pb/238U ages. (d) Tera-Wasserburg diagram for all analyses,

1642 intercepting the Concordia at 15.1 ± 0.2 Ma. Discordant U/Pb ages suggest a slight common

1643 Pb contamination and the uncertainty of the 207Pb/235U ages due to the low 207Pb content of

1644 young monazites. In this study, only the 208Pb/232Th ages were considered (inset c).

1645 Mineral abbreviations are after Whitney and Evans (2010)

1646

1647 Figure 14: Large-scale structure and available time-constraints

1648 (a) Three-dimensions large-scale sketch of Ikaria depicting the relationships between the

1649 structural dome, the synkinematic intrusions and the major shear zones. (b) Time-chart

1650 compiling all geochronological constraints available for Ikaria. Note that both the

1651 metamorphic dome and the late intrusions share a common fast cooling from 11-10 Ma

1652 onward. U/Pb ages on zircon (Zrn) are from Bolhar et al. (2010); K/Ar ages on muscovite

1653 (Ms) and biotiteAccepted (Bt), Rb/Sr ages on muscovite and Manuscript FT ages on apatite (Ap) are from Altherr

1654 et al. (1982) and Kumerics et al. (2005). (U-Th)/He ages on apatite and FT ages on zircon are

1655 from Kumerics et al. (2005); K/Ar ages on hornblende (Hbl) are from Altherr et al. (1982).

1656 Mean closure temperature are from Harrison (1981), Steck and Hunziker (1994), Grove and

1657 Harrison (1996), Brandon et al. (1998), Ketcham et al. (1999), Farley (2000), Cherniak and

Page 68 of 86 1658 Watson (2001) and Harrison et al. (2009). K/Ar, Ar/Ar and Rb/Sr ages on metamorphic rocks

1659 could reflect deformation-assisted crystallization (Kumerics et al., 2005) or cooling (Altherr et

1660 al., 1982). Same ages on granites are interpreted as cooling ages (Altherr et al., 1982;

1661 Kumerics et al., 2005). The new age of partial melting from this study is also showed.

1662 Question mark is associated with ages which are discussed in the text.

1663

1664 Figure 15: Large-scale implications

1665 (a) Tectonic map of synorogenic and postorogenic structures in the Aegean domain showing

1666 i) how far the back-arc postorogenic extension remains highly asymmetric in the center of the

1667 domain and ii) the footprint of the central Aegean HT zone where MCCs are described. Red

1668 and green arrows indicate the ductile sense of shear associated to the Oligo-Miocene

1669 extensional episode associated with the NCDS, the WCDS and the NDP (Lister et al., 1984;

1670 Faure et al., 1991; Gautier and Brun, 1994; Vandenberg and Lister, 1996; Jolivet and Patriat,

1671 1999; Mehl et al., 2005; 2007; Huet et al., 2009; Grasemann et al., 2012; Augier et al, 2015).

1672 On Syros and Sifnos islands, the synorogenic top-to-the-ENE sense of shear is associated to

1673 the synorogenic Vari detachment of Syros (Trotet et al., 2001a; 2001b). (b) Stretching-parallel

1674 cross-section through the central Aegean HT zone where intensity of stretching and

1675 asymmetry of the deformation are maximum.

Accepted Manuscript

Page 69 of 86 Figure1 Click here to download high resolution image

Accepted Manuscript

Page 70 of 86 Figure2 Click here to download high resolution image

Accepted Manuscript

Page 71 of 86 Figure3 Click here to download high resolution image

Accepted Manuscript

Page 72 of 86 Figure4 Click here to download high resolution image

Accepted Manuscript

Page 73 of 86 Figure5 Click here to download high resolution image

Accepted Manuscript

Page 74 of 86 Figure6 Click here to download high resolution image

Accepted Manuscript

Page 75 of 86 Figure7 Click here to download high resolution image

Accepted Manuscript

Page 76 of 86 Figure8 Click here to download high resolution image

Accepted Manuscript

Page 77 of 86 Figure9 Click here to download high resolution image

Accepted Manuscript

Page 78 of 86 Figure10 Click here to download high resolution image

Accepted Manuscript

Page 79 of 86 Figure11 Click here to download high resolution image

Accepted Manuscript

Page 80 of 86 Figure12 Click here to download high resolution image

Accepted Manuscript

Page 81 of 86 Figure13 Click here to download high resolution image

Accepted Manuscript

Page 82 of 86 Figure14 Click here to download high resolution image

Accepted Manuscript

Page 83 of 86 Figure15 Click here to download high resolution image

Accepted Manuscript

Page 84 of 86 Table1

Sample Coordinates (UTM 35 N) n R2 T (°C) Mean SD Mean SD Agios-Kirykos unit IK 12-15 441548 4168007 21 0,562 0,036 391 16 IK 13-33 438478 4163867 19 0,535 0,043 403 19 IK 13-31 438178 4163625 20 0,482 0,049 426 21 IK 13-34 438453 4164338 19 0,473 0,044 430 19 IK 12-13 439086 4164923 19 0,453 0,037 439 17 IK 13-22 440733 4168835 14 0,433 0,023 448 10

Ikaria unit IK 13-28 441898 4171296 19 0,312 0,049 502 22 IK 12-14 439237 4165283 20 0,303 0,082 506 36 IKS 13-13 428570 4164441 17 0,278 0,035 517 16 IKS 13-11 431485 4165492 21 0,269 0,104 521 46 IK 12-01 428758 4164611 23 0,266 0,051 522 22 IK 12-20 427111 4163553 18 0,264 0,038 523 17 IK 13-29 442324 4170727 13 0,262 0,034 524 15 IK 13-35 438248 4164495 15 0,251 0,039 530 17 IK 13-27 441743 4170619 15 0,241 0,033 534 15 IK 13-07 426986 4160696 21 0,233 0,081 537 36 IKS 13-06 432513 4162583 16 0,228 0,029 540 13 IKS 13-03 424903 4158649 16 0,219 0,047 543 21 IK 12-12 438723 4164980 20 0,216 0,082 545 37 IK 13-05 426436 4160041 15 0,216 0,055 545 24 IK 13-06 426795 4160494 16 0,213 0,065 546 29 IKS 13-12 427145 4161730 16 0,209 0,035 548 16 IK 12-04 426177 4165286 16 0,204 0,040 550 18 IK 13-08 425761 4165301 18 0,202 0,057 551 25 IK 12-21 425535 4161591 13 0,194 0,044 555 19 IK 12-11 438363 4165292 19 0,174 0,041 564 18 IKS 13-10 435557 4167534 13 0,130 0,042 583 19 IK 13-09 425573 4165499 17 0,129 0,045 584 20 IKS 13-04 430168 4161819 11 0,129 0,041 584 18 IKS 13-01 429434 4159919 17 0,110 0,051 592 23 IK 13-16 440281 4168893 18 0,107 0,051 594 22 IK 13-13 439031 4170692 15 0,083 0,040 604 18 IKS 13-14 427196 4158256 13 0,072 0,041 609 18 IKS 13-15 431859 4160264 14 0,054 0,045 617 20 IK 12-24 433322Accepted 4166192 18 0,037 0,034Manuscript625 15

Page 85 of 86 Supplementary data (e-component) Click here to download Supplementary data (e-component): Th-U-Pb analyses.xls

Accepted Manuscript

Page 86 of 86